Next Article in Journal
Alginate-Based Hydrogels and Scaffolds for Biomedical Applications
Previous Article in Journal
Effect of Storage Time and Temperature on the Bioactivity of a Chitosan-Derived Epigenetic Modulation Scaffold
Previous Article in Special Issue
Diversity and Distribution of Carotenogenic Algae in Europe: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Astaxanthin as a King of Ketocarotenoids: Structure, Synthesis, Accumulation, Bioavailability and Antioxidant Properties

1
Diatom Nanoengineering and Metabolism Laboratory (DNM), School of Applied Science, Dr. Harisingh Gour Central University, Sagar 470003, India
2
Department of Biotechnology, School of Engineering, Maharishi Markandeshwar University, Ambala 133203, India
3
State Forensic Science Laboratory, Madhuban 132037, India
4
Department of Environmental Engineering, College of Ocean Science and Engineering, Korea Maritim and Ocean University, Busan 49112, Republic of Korea
5
Department of Botany, Plant Biotechnology Laboratory, Mohanlal Sukhadia University, Udaipur 313001, India
6
School of Energy and Environment, City University of Hong Kong, Hong Kong 999077, China
7
Sustainability Cluster, School of Engineering, University of Petroleum and Energy Studies, Dehradun 248007, India
8
Metabolism, Bioengineering of Microalgal Metabolism and Applications (MIMMA), Biology of Organisms, Stress, Health and Environment, Le Mans University, IUML—FR 3473 CNRS, 72000 Le Mans, France
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Mar. Drugs 2023, 21(3), 176; https://doi.org/10.3390/md21030176
Submission received: 30 September 2022 / Revised: 7 March 2023 / Accepted: 8 March 2023 / Published: 13 March 2023
(This article belongs to the Special Issue Microalgal Carotenoids)

Abstract

:
Astaxanthin (3,3-dihydroxy-β, β-carotene-4,4-dione) is a ketocarotenoid synthesized by Haematococcus pluvialis/lacustris, Chromochloris zofingiensis, Chlorococcum, Bracteacoccus aggregatus, Coelastrella rubescence, Phaffia rhodozyma, some bacteria (Paracoccus carotinifaciens), yeasts, and lobsters, among others However, it is majorly synthesized by Haematococcus lacustris alone (about 4%). The richness of natural astaxanthin over synthetic astaxanthin has drawn the attention of industrialists to cultivate and extract it via two stage cultivation process. However, the cultivation in photobioreactors is expensive, and converting it in soluble form so that it can be easily assimilated by our digestive system requires downstream processing techniques which are not cost-effective. This has made the cost of astaxanthin expensive, prompting pharmaceutical and nutraceutical companies to switch over to synthetic astaxanthin. This review discusses the chemical character of astaxanthin, more inexpensive cultivating techniques, and its bioavailability. Additionally, the antioxidant character of this microalgal product against many diseases is discussed, which can make this natural compound an excellent drug to minimize inflammation and its consequences.

1. Introduction

Among existing carotenoids, about 600 different types are biosynthesized in plants, mosses, algae, bacteria, and fungi [1,2,3]. Most carotenoids are composed by 40 carbon atoms joined by double bonds (Figure 1) [4,5,6]. These have further been grouped into carotenes and xanthophylls [7,8,9]. Carotenes are composed of only carbon and hydrogen atoms, whereas xanthophylls are oxygenated derivatives which may have hydroxy groups (e.g., zeaxanthin), keto groups (e.g., canthaxanthin) or combinations of groups (e.g., astaxanthin, fucoxanthin, violaxanthin) [10,11,12].
Among carotenoids, the ketocarotenoid astaxanthin (3,3′-dihydroxy-β, β′-carotene-4,4′-dione) occupies a particular place because of its intense red color and antioxidant properties that are ca. 20–50 times higher than synthetic astaxanthin [13], but the antioxidant properties of free and esterified astaxanthin are still debated [14]. These unique properties of natural astaxanthin are among the primary reasons of its high value in the market which ranges from $6000 to $7150 per kg and the commercial market for astaxanthin has shown a great potential and is estimated to grow up to 3.4 billion by 2027. Nevertheless, cheap chemical astaxanthin meets 95% of the market demand [15,16,17]. Considering this price difference, replacing synthetic astaxanthin by natural astaxanthin is not cost effective and therefore requires strategies and techniques to increase astaxanthin production for cutting down the cost of its production. This includes photobioreactors and economically efficient downstream processing [18] and/or reconsidering astaxanthin sources besides Haematococcus lacustris, which is the major producer of astaxanthin (about 4% of dry weight) and is already cultivated at an industrial scale [19,20,21,22].
Astaxanthin was discovered in lobsters (Astacus gammarus) by the Nobel laureate Richard Kuhn [23]. Since this pioneer work, astaxanthin has been found to be naturally synthesized by land plants (Adonis annua: [24] and Adonis aestivalis: Egger and Kleinig [25]), microalgae (Haematococcus laccustris, Chromochloris zofingiensis (formerly Chlorella zofingiensis: Fucikova and Lewis [26], Chlorococcum sp.: Kopecky, et al. [27], Bracteacoccus aggregatus [28], Coelastrella rubescence [29], Chlamydomonas nivalis: Viala [30], yeasts (Phaffia rhodozyma) and some bacteria eg. Paracoccus carotinifaciens, a motile aerobic Gram-negative bacterium [31,32].
Astaxanthin, as with other carotenoids, is characterized by a polyene system allowing the molecules to exist in cis- and trans-isomeric forms. The cis-configuration is comparatively less stable than the trans-isomeric form [6]. Most of the carotenoids found in nature exhibit a trans-configuration. In addition, astaxanthin has chiral carbon at the C3 and C3′ positions, allowing the existence of two enantiomers (3R, 3′R and 3S, 3′S) and one meso-configuration (3R, 3′S) (Figure 2). Another type of chemical modification of astaxanthin is esterification with one or two fatty acids [33]. Altogether, these chemical particularities confer to the carotenoids typical molecular configurations that may impact their light absorbing properties [34], chemical stability and bioavailability [35]. For instance, mono-esterified astaxanthin with unsaturated fatty acids and a short chain is rapidly hydrolysed in the human digestive system [36], indicating that free astaxanthin bioavailability is higher in comparison to mono-esterified and diesterified astaxanthin.
Astaxanthin from Haematococcus lacustris is mainly the 3S, 3′S isomer which is 70% monoesterified, 25% diesterified and 5% free form with 3′-OH group modification [37], whereas that from Phaffia rhodozyma is 3R, 3′R unesterified astaxanthin due to its biosynthesis [38]. Astaxanthin from Paracocuss carotinifacience is 100% free form with 3,3′-OH group modification [39]. Note that synthetic astaxanthin is produced from petrochemical products [40] via a multistep process that includes hydroxylation of canthaxanthin, oxidation of zeaxanthin and the Wittig reaction [36] (Figure 3). Synthetic astaxanthin is a mixture of stereoisomers, 1(3R, 3′R), 2(3R, 3′S), and 1(3S, 3′S) nonesterified forms [41,42].
Only natural astaxanthin is considered safe for human use and consumption. Due to its multiple applications in pharmaceutical, nutraceutical, cosmetics, dietary and aquaculture, astaxanthin is produced naturally by many companies due to its rapidly growing list of attributes [43]. Chemically synthesized astaxanthin is not approved for direct consumption because of its chemical residues, which are not appropriate for humans [44] although they are recommended for aquaculture [45].
This review discusses the chemical structure and synthesis of carotenoids, mainly astaxanthin. It also describes cultivation strategies, primarily of Haematococcus lacustris for astaxanthin production. The review throws light on accumulation, harvesting, and bioavailability of astaxanthin. The pathophysiology of diseases caused by poor immunity, and the role of astaxanthin in regulating the immune responses is described. Additionally, a special focus is on the ability of astaxanthin to reduce inflammation and reactive oxygen species (ROS), resulting in several other health benefits.

2. Cultivation and Accumulation of Astaxanthin in Haematococcus lacustris

Haematococcus lacustris has three different life stages, i.e., macrozooids, microzooids, coccoid vegetative cells and palmella or haematocyst. Green macrozooids are found in a nonstressful environments, i.e., when all the required nutrients are available for multiplication, and is mostly asexual [46]. When one environmental condition becomes limiting, macrozooids are transformed to palmella and start to accumulate astaxanthin in the cytoplasm, requiring deep metabolic reorientation. At the same time, cells stop diving, enlarge and enter the red cyst stage, enriched in astaxanthin [47]. Figure 4 shows the morphological, biochemical and gene pathway stages for astaxanthin synthesis, and the mechanism [17,48]. From the carotenoid point of view, green macrozooids contain the typical carotenoids of green microalgae, i.e., β-carotene, lutein, violaxanthin and neoxanthin, whereas at cyst stage they have 90% astaxanthin, along with its mono and di esters [49].

2.1. Biochemical Pathway

In green microalgae the keto group in carotenoids is at the fourth position of the ionone rings. In contrast to β-carotene that accumulates in the plastoglobuli in the Dunaliella bardawil chloroplast matrix [50], astaxanthin accumulates in cytoplasmic lipid globules of Haematococcus lacustris [6].
In the biosynthetic pathway for astaxanthin synthesis (Figure 4), Isopentenyl pyrophosphate (IPP) is the precursor molecule [51] and can be produced via two pathways, i.e., the mevalonate (MVA) and methyl erythritol 4-phosphate (MEP) pathways. However, in Haematococcus it is produced via the MEP pathway only [52]. IPP is coded by two enzymes ipi1 and ipi2. IPP undergoes a set of reactions involving isomerization, condensation and desaturation yielding β-carotene. The β-carotene is then transferred from chloroplast to cytoplasm for the formation of astaxanthin in the presence of enzyme Cyt P450- β-carotene hydroxylase (CRTR-b), whose ketolase activity is present only in the cytoplasm, resulting in upregulation of crtr-b gene [53]. This is accompanied by an active tricarboxylic acid (TCA) cycle [54,55].

2.2. Wastewater as Growth Medium for Haematococcus

In the recent past, several companies have taken steps in the direction of production of natural astaxanthin, but the technology is still in its infancy with regard to bringing down the cost of production [56]. Although various media recipes are recommended, such as Bold Basal Medium (BBM), Blue Green medium (BG−11), the National Institute for Environmental Studies medium (NIES), Tris acetate phosphate (TAP) and the Optimal Haematococcus Medium (OHM) [57,58] for the growth of Haematococcus, none are economical for large-scale cultivation. On the offset, the cultivation can be made economical if Haematococcus lacustris is grown in pretreated wastewater since it has ability to remediate wastewater by utilizing the basic nutrients such as nitrates, phosphates and trace metals that are present in abundance in the wastewater [59]. The microalgae reduce the nitrates, phosphates, ammonia, sulphates and reduce the chemical oxygen demand (COD), while simultaneously producing biomass and various value added products [60,61]. Kang C.D et al., 2006 [62] found that Haematococcus grown in primary treated sewage and pretreated piggery wastewater diluted four times (filtered via 0.2 µm pores and pH set to 7.5) completely removed nitrates at 42.4 mgL−1 and 469 mgL−1,respectively, phosphates at 2.6 mgL−1 and 98.7 mgL−1, respectively, as well as other salts and COD. The astaxanthin accumulated in primary treated sewage and pretreated piggery wastewater was 39.7 mgL−1 and 83.9 mgL−1, respectively. In another study, pretreated potato wastewater was used as the growing nutrient medium for Haematococcus lacustris. It was found that astaxanthin accumulation in Haematococcus cells grown in potato wastewater reached to 24.5–27.9 mgg−1 in 3 days. This was much higher than for control cells, which showed only 14.7 mgg−1 astaxanthin accumulation in 12 days [63]. This is quite encouraging, as this simultaneously produced methane while remediating high strength organic wastewater with efficient removal of COD, phosphates, nitrates, and ammonia.

2.3. Growth of Haematococcus in Photosynthetic Microbial Fuel Cells

The latest technique for remediating wastewater is the use of microbial fuel cells [64]. Microbial fuel cells are biochemical devices in which electrons are produced by anaerobic bacteria during the process of degrading organic matter in wastewater at the anode. The redox reactions occurring at the anode and cathode are separated by a proton exchange membrane and connected to a resistor [65]. The process produces electricity while remediating wastewater pollutants at the anode. Further, if the electron acceptors at cathode are replaced by living microalgae, this is known as a photosynthetic microalgal microbial fuel cell [66,67]. Thus, if Haematococcus lacustris is used as a cathode substrate in a microbial fuel cell it produces free oxygen. Hence, in this process, the wastewater is remediated and bioelectricity is produced, besides the production of value-added products in the form of biomass and pigments [68]. Such kind of microbial fuel cell that hybridize with an algal photobioreactor separated by a proton exchange membrane such as a clay plate which are economical it can reduce the cost of astaxanthin production when upscaled [69,70,71]. Although in a microbial fuel cell, only electrogenic microbes are used at the anode, recently Khan et al. 2022 [72] found that diatoms (brown microalgae) are electrogenic when used at the anode due to NOX- ions produced in diatoms for iron reduction. The MFC thus produces electricity (12.6 mWm−2), lipids (22.31 μg mL−1 on 15th day) and pigments (4.890 μg mL−1 on 15th day). Further, the dead diatom frustules/biomass act as polyabsorbents for absorbing pollutants from the wastewater [73]. Similarly a schematic representation of photosynthetic MFC with Haematococcus at the cathode, and harvesting of its cells for astaxanthin can be seen in Figure 5.

2.4. Growth of Haematococcus in Plastic Bubble Wraps

The harvesting of Haematococcus lacustris for large scale production of astaxanthin is generally carried out in open raceways, i.e., closed photobioreactors at large scale, but it has its drawbacks. The open raceway often gets contaminated has irregular irradiation of light, and requires a regular supply of media and water, whereas closed photobioreactors requires expensive photobioreactors, a electricity supply, and maintenance of pH, CO2 and temperature, which is not economically viable [74]. However, the concept of bubble farming has been introduced in which algae are grown in plastic bubble wraps discarded from the plastics industry [75] (Figure 5II). In recent work on diatoms cultivated in different types of plastic bubble wraps, it has been established that low density polypropylene bubble wraps proved to be the best closed-chamber, air-filled photobioreactor, allowing the exchange of air but inhibiting water evaporation, and avoiding cell contamination. The cells grew exponentially and survived for months without losing their water content compared to cells kept in an open container. The chlorophyll content in bubble wraps reached to 3.34 μg mL−1 compared to 0.735 μg mL−1 in the control on 40th day and after [75]. The growth of Haematococcus was also tested in small scale plastic bubble wraps (Volume 160 µL in a bubble wrap of diameter 0.25”) and showed a similar pattern of gaseous exchange, no contamination, and multiplication of cells with no loss of water [75]. In addition to bubble wraps, all closed plastic containers made up of low density polypropylene, or any other plastic material that allows exchange of gases, inhibits water loss, and is transparent to light, are suitable for cultivating microalgal cells for their value-added products [76].

2.5. Stress Factors to Induce Astaxanthin Production in Haematococcus

The effect of carotenogenesis depends upon environmental factors such as temperature and light [77]. Optimum temperature promotes growth of cells, while light enhances cell multiplication during the juvenile stage of cells, and high light favours astaxanthin production in Haematococcus lacustris [49].
Haematococcus can be cultivated in two stages, first cultivating for the green stage, and then shifting the cells towards stress environment to simulate the red stage for astaxanthin production. There are many factors which may result in increasing astaxanthin content in Haematococcus lacustris in vitro, which include both physical and chemical factors. There have been intense studies on various chemicals to increase astaxanthin yield, and the genes responsible for its upregulation have been monitored via transcriptomics. High light and nitrogen deficiency are the most common factors to induce the shift from green cell stage of Haematococcus lacustris towards the red cell stage for astaxanthin synthesis [78]. Further it is the stress environment that increases reactive oxygen species level in the cell environment of Haematococcus that is responsible for increasing the astaxanthin level [79]. In addition, gamma amino butyric acid (GABA) maintains the stress tolerance of the cells under a stress environment [80]. Haematococcus cells showed the highest biomass (1.65 g L−1), of astaxanthin (3.86 mg L−1 d−1) and lipids (ca. 55.11%) when treated with 0.25 mM GABA. Additionally phytohormones and chemicals enhance astaxanthin synthesis and accumulation [54]. In addition, polyamines, such as putrescine, spermidine and spermine play an important role in regulating stress. For instance, melatonin and putrescine are stress signal chemicals that are inexpensive and improve astaxanthin synthesis in Haematococcus. These chemicals can be an appropriate choice for reducing culture and cultivation costs for astaxanthin production. Thorough research of all the chemicals and factors that enhance astaxanthin production in Haematococcus, along with their transcriptomic data, is necessary to analyze the upregulated genes and downregulated genes as shown in Table 1. However, there are still many unexplored chemicals and factors for which transcriptomics gene studies are pending. In brief, to make the whole process of astaxanthin production and accumulation cost-effective, different factors influencing the accumulation of astaxanthin and genes responsible, with their complete transcriptomic data, need to be explored.

3. Astaxanthin Harvesting and Its Bioavailability

3.1. Harvesting of Haematococcus for Astaxanthin

The rigid cell wall of Haematococcus lacustris (~2.2 µm) is made up of algaenan, biopolymers and thick polysaccharides [96,97]. Since, cell disruption in downstream processes such as pulse electric field current, microwaves, sonication, enzyme treatment, chemical solvents, high pressure homogenizers and bead millers, are expensive at large scale, they tend to increase the cost of extraction of astaxanthin [98,99] (Figure 6I). However, if among these, an economically viable technique is optimized that can lead to harvesting of astaxanthin from the cells, astaxanthin can become an important ingredient in all affordable diets and supplements [100,101]. A recent study has shown that on applying a mechanical strain, Haematococcus cells when experienced shear stress, oozed astaxanthin in a microfluidic chamber [102]. A microfluidic chamber with 65 micropillars with six different heights was used for selecting the Haematococcus cells put under mechanical stress. The gap between the micropillars was 10 µm to allow the young, flagellated cells to flow through. The height of drain channel was kept at 5 µm to prevent it from draining off. Mechanical stress was applied by loading the cells in culture chambers of different sizes ranging from 5 µm to 70 µm. Analysis done by Raman spectroscopy, showed that the amount of astaxanthin accumulated in different culture chambers inducing mechanical stress was maximum in a chamber with a height of 15 µm. This is similar to earlier reports on diatoms where mechanical shear strain generated by resonance energy in a microfluidic channel caused diatoms to ooze oil without lysing [103]. Studies have also shown that nanotechnology employing nanoparticles and laser light irradiation can electroporate the cell wall of Haematococcus. It has been observed that α-quartz nanoplates (NPLs) in combination with an ethyl-3-methylimidazolium (Emim) ionic liquid helps in extracting astaxanthin from Haematococcus at room temperature [104]. It was observed that extraction by ionic liquid alone was much less effective than the combination of ionic liquid and α-quartz NPLs. It was observed that when 90-days-old Haematococcus cells were treated with ionic liquids such as thiocyanate, diethylphosphate, HSO4, and Cl ions, the extraction of astaxanthin was quite low at about 9.6–14.2%. However, when these ionic liquids were combined with α-quartz NPLs to treat 90-days-old Haematococcus cell, it was found that the extraction efficiency increased and was highest for the ionic liquid HSO4 (80.7%). In a study by Praveen Kumar et al., 2015 [105], a gold nanoscalpel (Au-NS) (300 nm thick and 1–3 µm of width) synthesized by a chemical vapor transport method, and mounted on a three dimensional piezoelectric stage incised single cell of Haematococcus kept in an agarose microwell, showed 2 times more accumulation of astaxanthin. This milking, or harvesting, by Au-NS on Haematococcus cells leads to regular harvesting of astaxanthin without cell lysis. Thus, different techniques such as electroporation, use of nanoparticles, mechanical shear strain and nanoscalpels, may be considered for harvesting astaxanthin from Haematococcus, as shown in Figure 6IIA–D and III. The fractured cells are followed by healing and reaccumulation, compared to a normal cell (Figure 6IV).

3.2. Astaxanthin Bioavailability

Astaxanthin is a super antioxidant molecule having significant biological functions and the ability to cross the blood, brain and retinal barriers [106]. This depends upon its absorption, transport, and its metabolism mechanism in humans. Among the 600 types of identified carotenoids, around 50 carotenoids occur naturally in the food, and only about 20 carotenoids have the capability to be absorbed in the animal intestine and further distributed to tissues [107]. The absorption of astaxanthin in the human digestive system is further dependent upon bile salts, pH, and glycine/taurine ratios. The uptake and absorbance of carotene is four-fold higher in the presence of bile salts and as a part of a fatty food meal. Briefly, astaxanthin is absorbed with fatty acids into the intestinal epithelium via a passive diffusion mechanism. Further, astaxanthin combines with secretions from the spleen, resulting in the formation of micelles in the intestine, which is partially engrossed by intestinal mucosal cells. These cells deliver the astaxanthin to the liver, and in the liver, it mixes with lipoproteins and gets transported to the body tissues by circulation. Additionally, one of feature of astaxanthin molecules is their insertion into the lipid bilayer without degrading the cell membrane structure, thereby shielding the redox state and functional integrity of mitochondria [108]. The bioavailability of astaxanthin can be estimated by its chemical nature, such as its degree of esterification, geometrical isomer form, and optical stereoisomer form. The oral availability of astaxanthin, however, depends on its time of consumption before and after meals in non-smokers and smokers. It was found that the bioavailability of astaxanthin was higher at 7.219 ± 3.118 µg h/l in the after-meal group, and 2.968 ± 959 µg h/l in the before meal-group in non-smokers, and higher than in smokers before (p < 0.05) and after meals (6.468 µg h/l) [109]. Nanoencapsulation of astaxanthin improves its bioavailability as well as its bioabsorption by dispersing its insoluble bioactives [110,111]. Figure 6V shows the better solubility of encapsulated nano emulsion of astaxanthin over free astaxanthin. It has found by Edelman R et al. [112] that delivery by potato protein (PP) astaxanthin nanoparticles enhanced the solubility and bioavailability of astaxanthin compared to astaxanthin alone. The encapsulated astaxanthin did not degrade at all in an in vitro simulated digestion model in humans, whereas unencapsulated astaxanthin degraded in 120 min, and at the end of digestion (240 min). The fact that PP astaxanthin did not degrade was due to extremely low pH (2.5) and the property of protease inhibitors in the PP, which inhibited trypsin and chymotrypsin degradation. Although there are many encapsulation techniques, since astaxanthin is thermolabile, encapsulation methods that avoid heating are preferred. Since astaxanthin has properties to block cell cycle progression leading to apoptosis, a new method of incorporating astaxanthin with poly(lactic-co-glycolic acid) (PLGA)-encapsulated astaxanthin and processed with ultrasonically treated broccoli-derived extracellular vesicles (BEV) to form astaxanthin@PLGA@BEV has shown better anti-cancer property against human colon cancer cells (HT-29) than in vivo astaxanthin [113]. Additionally, surface polymers such as arabic gum improve encapsulation efficiency, and spray drying, ionotropic gelation, nanoprecipitation, emulsification-evaporation have shown to be cost-effective ways of encapsulating astaxanthin [114,115,116,117,118].

4. Anti-Oxidative Properties of Astaxanthin versus Other Carotenoids

Carotenoids have a potential role against viral and bacterial diseases by blocking the cellular angiotensin converting enzyme (ACE2) receptor, regulating inflammation, and modulating the peroxisome proliferator activator response (PPARγ) expression [119]. They thus have anti-inflammatory properties that help in decreasing the oxidative stress checking the invasion of cytokine storm. An adequate dose (4–12 mg/day) of astaxanthin for 3 months serves against many common infections in healthy people (Lu et al., 2022). Astaxanthin is reported to have higher antioxidant activity compared to several other carotenoids such as lycopene, lutein, zeaxanthin, α-carotene and β-carotene in human studies [120]. Astaxanthin from Haematococcus lacustris fed to rats had increased antioxidant effects [121]. Astaxanthin is completely digested by lipoprotein lipase found on the surface of cells and excreted by liver. The safe dosage for human consumption of astaxanthin has been increased from 4 mg/day to 12 mg/day by the European Food Safety Authority and the US Food and Drug Administration respectively, provided the intake is within 30 days [122]. A study in ochratoxin (OTA)-induced lung injury in a mouse model showed that astaxanthin protected the organs from oxidative damage and inflammation via the nuclear factor (Nrf2/NF-κB) pathway, which is related to maintaining innate immunity [123]. Carotenoids also have an overall effect in reducing the serum concentration of C-reactive protein (CRP) and interleukin -6 (IL-6) produced in response to infections. Amongst available carotenoid supplements, astaxanthin exhibits the most promising anti-inflammatory effects in humans [119].
Astaxanthin is not a known antiviral compound, but it has shown properties to suppress viral infections due to its antioxidant and anti-inflammatory effects. It plays a key role in regulating the cytokine storm responsible for inflammation. Astaxanthin based drugs inhibit viruses and other microbes by stopping their replication, gene expression and multiplication in host cells. Research reports have shown that astaxanthin has anti-inflammation, antioxidative, autophagy and anti-apoptosis effects in stopping pathogenesis (Figure 7). This starts with a change in nuclear factor (NF-κB) and interferon regulatory 3 (IRF3) of the host cell nuclei. It further leads to changes in the activity of transcription factors that in macrophage polarization cause expression of C-reactive proteins (CRP), ferritin, D-dimers, IFN-1, and pro-inflammatory cytokines such as tumor necrosis factor (TNF-α), IL-6, and monocyte chemoattractant factor (MCP-1). The increase in influx of these cytokines causes increase in mast cells, neutrophils and macrophages, which cause disruption in first line of defense against the virus leading to inflammation cytokine storm syndrome (CSS) and macrophage activation syndrome (MAS), as well as the disruption of lung function [124,125]. Astaxanthin has the capability to reduce oxidative stress and free radicals 65 times more effectively than vitamin C, 54 times more than carotene, and about 100 times more than tocopherol [126]. It has further shown inhibitory action on macrophage activation, phosphorylation of nuclear factor kappa B (NF-KB), Janus kinase, signal transducer and activator transcription (STAT), interleukin (IL)-6, IL-1β, cyclooxygenase- 2 and TNF-α factors [31,127,128,129].
The radical scavenging effect of α-tocopherol a fat-soluble antioxidant, is taken as a standard in comparison with β-carotene, which has lesser antioxidant properties. However, on comparing the antioxidant properties of canthaxanthin and astaxanthin in comparison with α-tocopherol and β-carotene investigated in a rat live microsome membrane exposed to 25 mMAAPH, 2,2′-azobis (2-amidinopropane); chelated Fe3+ and NADPH, induced free radical chain oxidation and formation of hydroperoxide formation (malondialdehyde (MDA)), it was observed that 2.3 nmol/mg protein of astaxanthin inhibited MDA and its hydroxide formation, whereas astaxanthin and canthaxanthin together inhibited at a rate greater than α-tocopherol alone. On the other hand, carotene at 10 nmol/mg protein was ineffective in inhibiting the MDA formation [130].
Astaxanthin is known to display greater biological activity compared to other anti-oxidants due to its ability to bind within the lipid bilayer of the cell membrane [131]. The polyene chain structure of astaxanthin traps the radicals inside the membrane of a cell, while the terminal ring of astaxanthin eliminates radicals from within the cell membrane as well as from its surface (Jinu & Mohan Chandra, 2021), as seen in Figure 7.
Furthermore, astaxanthin is reported to be upregulated in the irradiated cells of glutathione peroxidase 1, catalase, antioxidative enzymes, Nrf2 targeted proteins and heme oxygenase-1 (HO-1) (Jinu & Mohan Chandra, 2021). It blocks formation of ROS and triggers the expression of oxidative stress-responsive enzymes, including HO-1, which is further regulated by transcription factors such as Nrf2, reported to be activated by astaxanthin, thus protecting against oxidative stress in mice (Kubo et al., 2019).
Increased levels of ROS, as partially reduced oxygen metabolites, is strongly correlated with inflammation, oxidative injury, and viral infection and replication. It is speculated that regulating ROS level in virus-infected patients could be effective against hyperinflammation, reduce exacerbation of the immune system and protect tissues from oxidative injury, along with suppression of viral replication (Qin et al., 2020). The chemical properties and the molecular structure of astaxanthin explains its greater antioxidant activity. The natural form of astaxanthin with the all-trans isomer has better antioxidative properties than other carotenoids such as β-carotene, zeaxanthin, canthaxanthin, vitamin C, and vitamin E. Previous in vitro and in vivo studies on all-trans natural astaxanthin isomers have demonstrated its antioxidant effect and strong inhibition of lipid peroxidation (Liu & Osawa, 2007). In a study conducted by Liu et.al [132] in human neuroblastoma SH-SY5Y cells, it was reported that the cis astaxanthin (9-cis isomer) elevated antioxidant activity compared to the all-trans isomer in vitro, demonstrated by α, α-diphenyl-β-picrylhydrazyl (DPPH) scavenging activity test.

4.1. Immune System and Inflammation

Astaxanthin plays a crucial role in boosting the immune response [133], and is a multi-target pharmacological carotenoid that helps in treating neurological disorders such as Parkinson’s disease (PD), Alzheimer’s disease, depression, aging, and brain and spinal cord injuries [134]. It is also reported to have anti-inflammatory effects in several conditions such as neurodegenerative disorders, gastrointestinal disease, diabetes, and renal inflammation, which makes it a potential anti-inflammatory agent [135]. It also helps in T helpers 1 cytokines production, such as IL-2 and Mouse interferon Y (IFN-Y) (Fakhri et al., 2018). It also reduces nuclear factor-κB (NF-κB) and other down-stream mediators such as IL-1β, interleukin (IL)-6, matrix metalloproteinase (MMP-9). In addition, it modulates phosphoinositide 3-kinases (PI3K)/Akt, ERK/MAPK, and the up-stream macrophage migration inhibitory factor (MIF) [136].
The cytokine storm is a serious immune dysregulation caused by cytokine overproduction, which often occurs during virus infection, organ transplant, autoimmune diseases, and immunotherapy, and it may ultimately result in death if untreated (Qin et al., 2020). Along with destroying the invaded pathogen (mostly virus), the dysfunctional immune response results in diffusive alveolar damage and pulmonary oedema (Tay et al., 2020). High levels of inflammatory factors such as tumor necrosis factor-alpha (TNF—α) and interleukin-6 (IL-6), cause respiratory distress. These inflammatory factors can be controlled with the help of astaxanthin. Astaxanthin regulates the cyclooxygenase-2 (COX-2) pathway and suppress cytokines and other inflammatory agents (IL-6 and TNF—α).

4.2. Treatment of Skin Conditions and Protection of Eyes Health

Exposure of skin to ultra-violet radiation increases the risk of ROS level in the skin resulting in increased oxidative stress that triggers multiple oxidative actions such as damage to DNA. This can damage the skin and its physiological functions [137]. UV radiation induces skin cancer by initiating a chain reaction resulting in the generation of peroxides and other free radicles from lipids, and damaging DNA, causing the development of cancer cells [121]. Studies have shown that oral administration and absorption of astaxanthin, and its subsequent accumulation in tissues, protect the skin from radiations, aging and skin diseases. Different chemical forms of astaxanthin are responsible for its function. 13-cis astaxanthin had greater accumulation in the skin and eyes of rodents than the all-trans and 9-cis isomers after two weeks of continuous administration (10 mg/mL) [138]. Moreover, studies have reported that if astaxanthin is included in a regular diet it has the potential to enhance skin quality, increase skin elasticity, and reduce facial wrinkles and pigmentation [139].
Even though the function of the eyes is the absorption of light, long-term exposure of light may cause oxidative stress causing destructive structural and functional changes in the lens, retina and eye tissues [140]. Moreover, long-term exposure of light to eyes may induce and increase ROS level, which can activate cellular pathways associated with inflammation. It has been demonstrated that astaxanthin can cross the blood-brain barrier and accumulate in the retina of animals [141]. Therefore, the consumption of astaxanthin can help maintain eye health by reducing oxidative inflammation and asthenopia by improving retinal blood flow. A study conducted by Harada et al., [142] found that corneal tissues of mice that were affected with photokeratitis were healed, and inflammation was decreased, when astaxanthin nano powder was given orally. This study suggested that the inflammation reduction was due to reduction in the expression of cyclocxygenase-2 (COX-2), phosphorylated inhibitor of κB-α (p-IκB-α), tumor necrosis factor-α (TNF-α), and CD45 in the corneal tissue. Astaxanthin can enhance eye blood circulation and help in curing the eyestrain [139,143].

5. Safety Measures to Fight against Infectious Diseases Using Astaxanthin

The broad range of therapeutic advantages of astaxanthin make it a potential agent in the regulation of the host immune system. Numerous studies have revealed the potential of astaxanthin in healthcare and in fighting certain diseases as an effective antioxidant, anti-inflammatory agent, immune system modulator and glucose-level lowering agent. Studies have reported that astaxanthin supplementation lowers oxidative biomarkers such as MDA and isoprostanes and increased antioxidant capability, which helped in decreasing low density lipoprotein (LDL) oxidation and alleviated lipid peroxidation [111,128,129]. These supplements are also known to reduce the plasma levels of 12 and 15 hydroxy fatty acids in males in about three months [144]. Astaxanthin supplementation modulates the JAK (Janus Kinase)/STAT3 (Signal Transducers and Activators of Transcription) signaling pathway. JAK kinases are a family of tyrosine kinases that mediate the signaling by tyrosine phosphorylation of various proteins and receptor chains bound and dimerized by STAT for their translocation to the nucleus for activation or inactivation of transcription. Astaxanthin also regulates the human immune system by stimulating mitogen-induced lymphocyte proliferation, delaying hypersensitivity, and increasing levels of natural killer cell cytotoxicity and the total count of T-cells and B-cells in the peripheral blood [133]. Detailed studies of astaxanthin on different animal models targeting specific genes for therapeutic study is listed in Table 2. Astaxanthin is also reported to be an IgM and IgG stimulant in mouse spleen cells and human immunoglobulins in human cells. Studies in various animal models have demonstrated the ability of astaxanthin to activate the forkhead box O3 gene (FOXO3) that controls cell fate, metabolism, resistance to DNA damage, autophagy and apoptosis [145,146]. It is also responsible for regulating aging and age-related diseases, such as cardiovascular diseases, diabetes, cancers and other neurodegenerative diseases [145].

6. Future Perspectives of Astaxanthin as a Novel Drug

Research and development processes for new drugs represent a continuous fight against infectious diseases. Epidemic situations create pressure in the search for new drugs that do not generate side effects [171]. Many proof-of-concept studies have been undertaken, but controlled trials are necessary to establish the effectiveness of astaxanthin in the treatment of numerous diseases [172]. Currently, researchers are conducting clinical trials, investigating the remarkable effects of astaxanthin on inflammation and oxidative stress.
In the future, astaxanthin may play a role as significant as statins in the treatment and possible prevention of cardiovascular, pulmonary and neurogenerative diseases [173], although its popularity is limited due to its instability at high temperatures, under low pH and illuminated conditions, its hydrophobic nature, and limited bioavailability [111]. Astaxanthin applications are limited in biomedical applications due to its unstable nature, but methods such as encapsulation can be useful to preserve the biochemical activity of this biomolecule for longer periods. A lot of research has been carried out on its biochemical, biophysical, and molecular properties, but its transportation, metabolism and interaction studies need to be explored in detail [174]. Astaxanthin has great health benefits against many diseases when obtained from natural sources. Innovation towards the development of biocompatible, low-cost extraction methods is required for sustainable use of astaxanthin in the nutraceutical sector.

7. Conclusions

Astaxanthin is a super ketocarotenoid with unique chemical isomerization in different source organisms. However, astaxanthin from Haematococcus lacustris is most widely cultivated because of its safe and strong antioxidant nature. Haematococcus lacustris can be mass cultivated and to produce a novel drug and dietary supplement if the cost of its production can be reduced using different techniques and strategies. Further the bioavailability of astaxanthin dosage is an important factor that requires a deeper study and extensive research if it is multifunctional applications in different pharmaceutical, nutraceutical, cosmetic and dietary industries are be considered. Development of innovative drugs such as natural astaxanthin from microorganisms, and especially from Haematococcus lacustris, can lead to solutions for enormous scientific, clinical, and societal problems. Outbreaks of infectious diseases have created major public health emergencies globally. Sustained and consistent research is required to enhance our knowledge of key aspects of viral pathogenesis that can lead to improved precautionary and therapeutic policies.

Author Contributions

A.N., A.A., S.S., R.L., A.L., A.R. and D.A.J.: Literature review, Writing—original draft, Data curation; H.: Review & editing; S.V.: Supervision; Writing—original draft, Review & editing; G.S.: Figure 1, Figure 2, Figure 3 and review and editing, J.M.: Review & editing; B.S.: Review & editing; V.V.: Conceptualization, Supervision; Writing—original draft, Review & editing, Funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

Indo-French Centre for the Promotion of Advanced Research (IFCPAR/CEFIPRA) project number (PPMB-7133/2020) Indo France project.

Data Availability Statement

Not applicable.

Acknowledgments

AA: would like to thank CEFIPRA Indo French project for the junior research Fellowship. All authors would like to thank Indo-French Centre for the Promotion of Advanced Research (IFCPAR/CEFIPRA) project number (PPMB-7133/2020) Indo France project sanctioned to VV, H and BS for the financial aids.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schoefs, B.T. Chlorophyll and carotenoid analysis in food products. Properties of the pigments and methods of analysis. Trends Food Sci. Technol. 2002, 13, 361–371. [Google Scholar] [CrossRef]
  2. Takaichi, S. Carotenoids in phototrophic microalgae: Distributions and biosynthesis. In Pigments from Microalgae Handbook; Springer: Berlin/Heidelberg, Germany, 2020; pp. 19–41. [Google Scholar]
  3. Pagano, M.; Corrêa, E.; Duarte, N.; Yelikbayev, B. Pharmacological Potential of Pigments. In Biomolecules from Natural Sources: Advances and Applications; John Wiley & Sons: Hoboken, NJ, USA, 2022; pp. 101–112. [Google Scholar]
  4. Britton, G. Structure and properties of carotenoids in relation to function. FASEB J. 1995, 9, 1551–1558. [Google Scholar] [CrossRef] [PubMed]
  5. Christaki, E.; Bonos, E.; Giannenas, I.; Florou-Paneri, P. Functional properties of carotenoids originating from algae. J. Sci. Food Agric. 2013, 93, 5–11. [Google Scholar] [CrossRef] [PubMed]
  6. Lemoine, Y.; Schoefs, B. Secondary ketocarotenoid astaxanthin biosynthesis in algae: A multifunctional response to stress. Photosynth. Res. 2010, 106, 155–177. [Google Scholar] [CrossRef]
  7. Maoka, T. Carotenoids as natural functional pigments. J. Nat. Med. 2020, 74, 1–16. [Google Scholar] [CrossRef] [Green Version]
  8. Haugan, J.A. Algal carotenoids 54. Carotenoids of brown algae (Phaeophyceae). Biochem. Syst. Ecol. 1994, 22, 31–41. [Google Scholar] [CrossRef]
  9. Goodwin, T.W. Carotenoids and reproduction. Biol. Rev. 1950, 25, 391–413. [Google Scholar] [CrossRef]
  10. Riaz, M.; Zia-Ul-Haq, M.; Dou, D. Chemistry of carotenoids. In Carotenoids: Structure and Function in the Human Body; Springer: Berlin/Heidelberg, Germany, 2021; pp. 43–76. [Google Scholar]
  11. Mordi, R.C.; Ademosun, O.T.; Ajanaku, C.O.; Olanrewaju, I.O.; Walton, J.C. Free radical mediated oxidative degradation of carotenes and xanthophylls. Molecules 2020, 25, 1038. [Google Scholar] [CrossRef] [Green Version]
  12. Miller, N.J.; Sampson, J.; Candeias, L.P.; Bramley, P.M.; Rice-Evans, C.A. Antioxidant activities of carotenes and xanthophylls. FEBS Lett. 1996, 384, 240–242. [Google Scholar] [CrossRef] [Green Version]
  13. Ahirwar, A.; Kesharwani, K.; Deka, R.; Muthukumar, S.; Khan, M.J.; Rai, A.; Vinayak, V.; Varjani, S.; Joshi, K.B.; Morjaria, S. Microalgal drugs: A promising therapeutic reserve for the future. J. Biotechnol. 2022, 349, 32–46. [Google Scholar] [CrossRef]
  14. Régnier, P.; Bastias, J.; Rodriguez-Ruiz, V.; Caballero-Casero, N.; Caballo, C.; Sicilia, D.; Fuentes, A.; Maire, M.; Crepin, M.; Letourneur, D. Astaxanthin from Haematococcus pluvialis prevents oxidative stress on human endothelial cells without toxicity. Mar. Drugs 2015, 13, 2857–2874. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Scarsini, M.; Marchand, J.; Schoefs, B. Carotenoid overproduction in microalgae: Biochemical and genetic engineering. In Pigments from Microalgae Handbook; Jacob-Lopes, E., Queiroz, M.I., Zepka, L.Q., Eds.; Springer International Publishing: Cham, Switzerland, 2020; pp. 81–126. [Google Scholar] [CrossRef]
  16. Stachowiak, B.; Szulc, P. Astaxanthin for the food industry. Molecules 2021, 26, 2666. [Google Scholar] [CrossRef]
  17. Ahirwar, A.; Meignen, G.; Khan, M.J.; Sirotiya, V.; Scarsini, M.; Roux, S.; Marchand, J.; Schoefs, B.; Vinayak, V. Light modulates transcriptomic dynamics upregulating astaxanthin accumulation in Haematococcus: A review. Bioresour. Technol. 2021, 340, 125707. [Google Scholar] [CrossRef] [PubMed]
  18. Vinayak, V.; Manoylov, K.M.; Gateau, H.; Blanckaert, V.; Herault, J.; Pencreac’h, G.; Marchand, J.; Gordon, R.; Schoefs, B. Diatom milking: A review and new approaches. Mar. Drugs 2015, 13, 2629–2665. [Google Scholar] [CrossRef] [Green Version]
  19. Li, J.; Zhu, D.; Niu, J.; Shen, S.; Wang, G. An economic assessment of astaxanthin production by large scale cultivation of Haematococcus pluvialis. Biotechnol. Adv. 2011, 29, 568–574. [Google Scholar] [CrossRef]
  20. Olaizola, M. Commercial production of astaxanthin from Haematococcus pluvialis using 25,000-liter outdoor photobioreactors. J. Appl. Phycol. 2000, 12, 499–506. [Google Scholar] [CrossRef]
  21. Fei, Z.; Fan, F.; Liao, J.; Wan, M.; Bai, W.; Wang, W.; He, M.; Li, Y. Improving astaxanthin production of Haematococcus pluvialis on the outdoor large scale cultivation by optimizing the disinfection strategy of photobioreactor. Algal Res. 2022, 64, 102708. [Google Scholar] [CrossRef]
  22. Chekanov, K. Diversity and Distribution of Carotenogenic Algae in Europe: A Review. Mar. Drugs 2023, 21, 108. [Google Scholar] [CrossRef] [PubMed]
  23. Kuhn, R.; Sörensen, N.A. Über astaxanthin und ovoverdin. Ber. Der Dtsch. Chem. Ges. A B Ser. 1938, 71, 1879–1888. [Google Scholar] [CrossRef]
  24. Seybold, A.; Goodwin, T. Occurrence of astaxanthin in the flower petals of Adonis annua L. Nature 1959, 184, 1714–1715. [Google Scholar] [CrossRef]
  25. Egger, K.; Kleinig, H. Die ketocarotinoide in Adonis annua L.—II.: Zur struktur der ester. Phytochemistry 1967, 6, 437–440. [Google Scholar] [CrossRef]
  26. Fucikova, K.; Lewis, L.A. Intersection of Chlorella, Muriella and Bracteacoccus: Resurrecting the genus Chromochloris kol et chodat (Chlorophyceae, Chlorophyta). Fottea 2012, 12, 83–93. [Google Scholar] [CrossRef] [Green Version]
  27. Kopecky, J.; Schoefs, B.; Loest, K.; Stys, D.; Pulz, O. Microalgae as a source for secondary carotenoid production: A screening study. Algol. Stud. 2000, 98, 153–168. [Google Scholar] [CrossRef]
  28. Chekanov, K.; Litvinov, D.; Fedorenko, T.; Chivkunova, O.; Lobakova, E. Combined production of astaxanthin and β-carotene in a new strain of the microalga Bracteacoccus aggregatus BM5/15 (IPPAS C-2045) cultivated in photobioreactor. Biology 2021, 10, 643. [Google Scholar] [CrossRef]
  29. Ali, H.E.A.; Vorisek, F.; Dowd, S.E.; Kesner, S.; Song, Y.; Qian, D.; Crocker, M. Formation of Lutein, β-Carotene and Astaxanthin in a Coelastrella sp. Isolate. Molecules 2022, 27, 6950. [Google Scholar] [CrossRef]
  30. Viala, G. Recherches sur le Chlamydomonas nivalis Wille dans les Pyrénées. Bull. De La Société Bot. De Fr. 1967, 114, 75–79. [Google Scholar] [CrossRef]
  31. Yuan, J.P.; Peng, J.; Yin, K.; Wang, J.H. Potential health-promoting effects of astaxanthin: A high-value carotenoid mostly from microalgae. Mol. Nutr. Food Res. 2011, 55, 150–165. [Google Scholar] [CrossRef]
  32. Tsubokura, A.; Yoneda, H.; Mizuta, H. Paracoccus carotinifaciens sp. nov., a new aerobic gram-negative astaxanthin-producing bacterium. Int. J. Syst. Evol. Microbiol. 1999, 49, 277–282. [Google Scholar] [CrossRef] [Green Version]
  33. Lemoine, Y.; Rmiki, N.-E.; Créach, A.; Rachidi, J.B.S. Cytoplasmic accumulation of astaxanthin by the green alga Haematococcus pluvialis (Flotow). In Plant Cell Compartments; Schoefs, B., Ed.; Research Signpost: Kerala, India, 2008; pp. 251–284. [Google Scholar]
  34. WongCarter, K.; Llansola-Portoles, M.J.; Kodis, G.; Gust, D.; Moore, A.L.; Moore, T.A.; Croce, R.; van Grondelle, R.; van Amerongen, H.; van Stokkum, I. Light harvesting, photoregulation, and photoprotection in selected artificial photosynthetic systems. In Light Harvesting in Photosynthesis; CRC Press: Boca Raton, FL, USA, 2018; pp. 485–510. [Google Scholar]
  35. Qiao, X.; Yang, L.; Hu, X.; Cao, Y.; Li, Z.; Xu, J.; Xue, C. Characterization and evaluation of inclusion complexes between astaxanthin esters with different molecular structures and hydroxypropyl-β-cyclodextrin. Food Hydrocoll. 2021, 110, 106208. [Google Scholar] [CrossRef]
  36. Jafari, Z.; Bigham, A.; Sadeghi, S.; Dehdashti, S.M.; Rabiee, N.; Abedivash, A.; Bagherzadeh, M.; Nasseri, B.; Karimi-Maleh, H.; Sharifi, E. Nanotechnology-Abetted astaxanthin formulations in multimodel therapeutic and biomedical applications. J. Med. Chem. 2021, 65, 2–36. [Google Scholar] [CrossRef]
  37. Yang, L.; Qiao, X.; Gu, J.; Li, X.; Cao, Y.; Xu, J.; Xue, C. Influence of molecular structure of astaxanthin esters on their stability and bioavailability. Food Chem. 2021, 343, 128497. [Google Scholar] [CrossRef]
  38. Chen, Y.-Y.; Lee, P.-C.; Wu, Y.-L.; Liu, L.-Y. In vivo effects of free form astaxanthin powder on anti-oxidation and lipid metabolism with high-cholesterol diet. PLoS ONE 2015, 10, e0134733. [Google Scholar] [CrossRef] [PubMed]
  39. Sapone, V.; Iannone, A.; Alivernini, A.; Cicci, A.; Jessop, P.G.; Bravi, M. An innovative simplified one-pot process for Astaxanthin purification from Paracoccus carotinifaciens. Sep. Purif. Technol. 2023, 308, 122843. [Google Scholar] [CrossRef]
  40. Panis, G.; Carreon, J.R. Commercial astaxanthin production derived by green alga Haematococcus pluvialis: A microalgae process model and a techno-economic assessment all through production line. Algal Res. 2016, 18, 175–190. [Google Scholar] [CrossRef] [Green Version]
  41. Liu, X.; Luo, Q.; Cao, Y.; Goulette, T.; Liu, X.; Xiao, H. Mechanism of different stereoisomeric astaxanthin in resistance to oxidative stress in Caenorhabditis elegans. J. Food Sci. 2016, 81, H2280–H2287. [Google Scholar] [CrossRef] [PubMed]
  42. Sun, J.; Yan, J.; Dong, H.; Gao, K.; Yu, K.; He, C.; Mao, X. Astaxanthin with different configurations: Sources, activity, post-modification and application in foods. Curr. Opin. Food Sci. 2022, 49, 100955. [Google Scholar] [CrossRef]
  43. Gateau, H.; Solymosi, K.; Marchand, J.; Schoefs, B. Carotenoids of microalgae used in food industry and medicine. Mini Rev. Med. Chem. 2017, 17, 1140–1172. [Google Scholar] [CrossRef] [PubMed]
  44. Capelli, B.; Talbott, S.; Ding, L. Astaxanthin sources: Suitability for human health and nutrition. Funct. Foods Health Dis. 2019, 9, 430–445. [Google Scholar] [CrossRef] [Green Version]
  45. Capelli, B.; Bagchi, D.; Cysewski, G.R. Synthetic astaxanthin is significantly inferior to algal-based astaxanthin as an antioxidant and may not be suitable as a human nutraceutical supplement. Nutrafoods 2013, 12, 145–152. [Google Scholar] [CrossRef]
  46. Onorato, C.; Rösch, C. Comparative life cycle assessment of astaxanthin production with Haematococcus pluvialis in different photobioreactor technologies. Algal Res. 2020, 50, 102005. [Google Scholar] [CrossRef]
  47. Khoo, K.S.; Lee, S.Y.; Ooi, C.W.; Fu, X.; Miao, X.; Ling, T.C.; Show, P.L. Recent advances in biorefinery of astaxanthin from Haematococcus pluvialis. Bioresour. Technol. 2019, 288, 121606. [Google Scholar] [CrossRef] [PubMed]
  48. Zhao, Y.; Yue, C.; Geng, S.; Ning, D.; Ma, T.; Yu, X. Role of media composition in biomass and astaxanthin production of Haematococcus pluvialis under two-stage cultivation. Bioprocess Biosyst. Eng. 2019, 42, 593–602. [Google Scholar] [CrossRef]
  49. Schoefs, B.t.; Rmiki, N.-E.; Rachadi, J.; Lemoine, Y. Astaxanthin accumulation in Haematococcus requires a cytochrome P450 hydroxylase and an active synthesis of fatty acids. Febs Lett. 2001, 500, 125–128. [Google Scholar] [CrossRef] [Green Version]
  50. Xu, Y.; Ibrahim, I.M.; Wosu, C.I.; Ben-Amotz, A.; Harvey, P.J. Potential of new isolates of Dunaliella salina for natural β-carotene production. Biology 2018, 7, 14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Lichtenthaler, H.K. [34] Chlorophylls and carotenoids: Pigments of photosynthetic biomembranes. In Methods in Enzymology; Elsevier: Amsterdam, The Netherlands, 1987; Volume 148, pp. 350–382. [Google Scholar]
  52. Luo, Q.; Bian, C.; Tao, M.; Huang, Y.; Zheng, Y.; Lv, Y.; Li, J.; Wang, C.; You, X.; Jia, B. Genome and transcriptome sequencing of the astaxanthin-producing green microalga, Haematococcus pluvialis. Genome Biol. Evol. 2019, 11, 166–173. [Google Scholar] [CrossRef] [Green Version]
  53. Schoefs, B. The protochlorophyllide–chlorophyllide cycle. Photosynth. Res. 2001, 70, 257–271. [Google Scholar] [CrossRef] [PubMed]
  54. Cui, J.; Yu, C.; Zhong, D.-b.; Zhao, Y.; Yu, X. Melatonin and calcium act synergistically to enhance the coproduction of astaxanthin and lipids in Haematococcus pluvialis under nitrogen deficiency and high light conditions. Bioresour. Technol. 2020, 305, 123069. [Google Scholar] [CrossRef]
  55. Li, K.; Cheng, J.; Ye, Q.; He, Y.; Zhou, J.; Cen, K. In vivo kinetics of lipids and astaxanthin evolution in Haematococcus pluvialis mutant under 15% CO2 using Raman microspectroscopy. Bioresour. Technol. 2017, 244, 1439–1444. [Google Scholar] [CrossRef] [PubMed]
  56. Del Campo, J.A.; García-González, M.; Guerrero, M.G. Outdoor cultivation of microalgae for carotenoid production: Current state and perspectives. Appl. Microbiol. Biotechnol. 2007, 74, 1163–1174. [Google Scholar] [CrossRef] [PubMed]
  57. Waissman-Levy, N.; Leu, S.; Khozin-Goldberg, I.; Boussiba, S. Manipulation of trophic capacities in Haematococcus pluvialis enables low-light mediated growth on glucose and astaxanthin formation in the dark. Algal Res. 2019, 40, 101497. [Google Scholar] [CrossRef]
  58. Nahidian, B.; Ghanati, F.; Shahbazi, M.; Soltani, N. Effect of nutrients on the growth and physiological features of newly isolated Haematococcus pluvialis TMU1. Bioresour. Technol. 2018, 255, 229–237. [Google Scholar] [CrossRef]
  59. Kaloudas, D.; Pavlova, N.; Penchovsky, R. Phycoremediation of wastewater by microalgae: A review. Environ. Chem. Lett. 2021, 19, 2905–2920. [Google Scholar] [CrossRef]
  60. Ledda, C.; Tamiazzo, J.; Borin, M.; Adani, F. A simplified process of swine slurry treatment by primary filtration and Haematococcus pluvialis culture to produce low cost astaxanthin. Ecol. Eng. 2016, 90, 244–250. [Google Scholar] [CrossRef]
  61. Stevčić, Č.; Pulkkinen, K.; Pirhonen, J. Screening of microalgae and LED grow light spectra for effective removal of dissolved nutrients from cold-water recirculating aquaculture system (RAS) wastewater. Algal Res. 2019, 44, 101681. [Google Scholar] [CrossRef]
  62. Kang, C.D.; An, J.Y.; Park, T.H.; Sim, S.J. Astaxanthin biosynthesis from simultaneous N and P uptake by the green alga Haematococcus pluvialis in primary-treated wastewater. Biochem. Eng. J. 2006, 31, 234–238. [Google Scholar] [CrossRef]
  63. Pan, M.; Zhu, X.; Pan, G.; Angelidak, I. Integrated valorization system for simultaneous high strength organic wastewater treatment and astaxanthin production from Haematococcus pluvialis. Bioresour. Technol. 2021, 326, 124761. [Google Scholar] [CrossRef] [PubMed]
  64. Logan, B.E.; Hamelers, B.; Rozendal, R.; Schröder, U.; Keller, J.; Freguia, S.; Aelterman, P.; Verstraete, W.; Rabaey, K. Microbial fuel cells: Methodology and technology. Environ. Sci. Technol. 2006, 40, 5181–5192. [Google Scholar] [CrossRef]
  65. Logan, B.E. Exoelectrogenic bacteria that power microbial fuel cells. Nat. Rev. Microbiol. 2009, 7, 375–381. [Google Scholar] [CrossRef] [PubMed]
  66. Ahirwar, A.; Das, S.; Das, S.; Yang, Y.-H.; Bhatia, S.K.; Vinayak, V.; Ghangrekar, M.M. Photosynthetic microbial fuel cell for bioenergy and valuable production: A review of circular bio-economy approach. Algal Res. 2023, 70, 102973. [Google Scholar] [CrossRef]
  67. Vinayak, V.; Khan, M.J.; Varjani, S.; Saratale, G.D.; Saratale, R.G.; Bhatia, S.K. Microbial fuel cells for remediation of environmental pollutants and value addition: Special focus on coupling diatom microbial fuel cells with photocatalytic and photoelectric fuel cells. J. Biotechnol. 2021, 338, 5–19. [Google Scholar] [CrossRef] [PubMed]
  68. Khan, M.J.; Rai, A.; Ahirwar, A.; Sirotiya, V.; Mourya, M.; Mishra, S.; Schoefs, B.; Marchand, J.; Bhatia, S.K.; Varjani, S. Diatom microalgae as smart nanocontainers for biosensing wastewater pollutants: Recent trends and innovations. Bioengineered 2021, 12, 9531–9549. [Google Scholar] [CrossRef] [PubMed]
  69. Khan, M.J.; Singh, N.; Mishra, S.; Ahirwar, A.; Bast, F.; Varjani, S.; Schoefs, B.; Marchand, J.; Rajendran, K.; Banu, J.R. Impact of light on microalgal photosynthetic microbial fuel cells and removal of pollutants by nanoadsorbent biopolymers: Updates, challenges and innovations. Chemosphere 2022, 288, 132589. [Google Scholar] [CrossRef] [PubMed]
  70. Deka, R.; Shreya, S.; Mourya, M.; Sirotiya, V.; Rai, A.; Khan, M.J.; Ahirwar, A.; Schoefs, B.; Bilal, M.; Saratale, G.D. A techno-economic approach for eliminating dye pollutants from industrial effluent employing microalgae through microbial fuel cells: Barriers and perspectives. Environ. Res. 2022, 212, 113454. [Google Scholar] [CrossRef] [PubMed]
  71. Chatterjee, P.; Ghangrekar, M. Design of clayware separator-electrode assembly for treatment of wastewater in microbial fuel cells. Appl. Biochem. Biotechnol. 2014, 173, 378–390. [Google Scholar] [CrossRef]
  72. Khan, M.J.; Das, S.; Vinayak, V.; Pant, D.; Ghangrekar, M. Live diatoms as potential biocatalyst in a microbial fuel cell for harvesting continuous diafuel, carotenoids and bioelectricity. Chemosphere 2022, 291, 132841. [Google Scholar] [CrossRef]
  73. Rai, A.; Sirotiya, V.; Mourya, M.; Khan, M.J.; Ahirwar, A.; Sharma, A.K.; Kawatra, R.; Marchand, J.; Schoefs, B.; Varjani, S. Sustainable treatment of dye wastewater by recycling microalgal and diatom biogenic materials: Biorefinery perspectives. Chemosphere 2022, 305, 135371. [Google Scholar] [CrossRef]
  74. Sirohi, R.; Sim, S.J.; Pandey, A. Photobioreactors: An introduction. In Current Developments in Biotechnology and Bioengineering; Elsevier: Amsterdam, The Netherlands, 2023; pp. 3–10. [Google Scholar]
  75. Khan, M.J.; Gordon, R.; Varjani, S.; Vinayak, V. Employing newly developed plastic bubble wrap technique for biofuel production from diatoms cultivated in discarded plastic waste. Sci. Total Environ. 2022, 823, 153667. [Google Scholar] [CrossRef]
  76. Rai, I.; Ahirwar, A.; Rai, A.; Varjani, S.; Vinayak, V. Biowaste recycling strategies for regenerative life support system: An overview. Sustain. Energy Technol. Assess. 2022, 53, 102525. [Google Scholar] [CrossRef]
  77. Sirotiya, V.; Ahirwar, A.; Mourya, M.; Khan, M.J.; Rai, A.; Kwatra, R.; Sharma, A.K.; Schoefs, B.; Marchand, J.; Varjani, S. Astaxanthin bioaccumulation in microalgae under environmental stress simulated in industrial effluents highlighting prospects of Haematococcus pluvialis: Knowledge gaps and prospective approaches. Phytochem. Rev. 2022, 1–26. [Google Scholar] [CrossRef]
  78. Cui, D.; Hu, C.; Zou, Z.; Sun, X.; Shi, J.; Xu, N. Comparative transcriptome analysis unveils mechanisms underlying the promoting effect of potassium iodide on astaxanthin accumulation in Haematococcus pluvialis under high light stress. Aquaculture 2020, 525, 735279. [Google Scholar] [CrossRef]
  79. Zhao, Y.; Wang, H.-P.; Han, B.; Yu, X. Coupling of abiotic stresses and phytohormones for the production of lipids and high-value by-products by microalgae: A review. Bioresour. Technol. 2019, 274, 549–556. [Google Scholar] [CrossRef]
  80. Li, Q.; Zhao, Y.; Ding, W.; Han, B.; Geng, S.; Ning, D.; Ma, T.; Yu, X. Gamma-aminobutyric acid facilitates the simultaneous production of biomass, astaxanthin and lipids in Haematococcus pluvialis under salinity and high-light stress conditions. Bioresour. Technol. 2021, 320, 124418. [Google Scholar] [CrossRef] [PubMed]
  81. Gao, Z.; Li, Y.; Wu, G.; Li, G.; Sun, H.; Deng, S.; Shen, Y.; Chen, G.; Zhang, R.; Meng, C. Transcriptome analysis in Haematococcus pluvialis: Astaxanthin induction by salicylic acid (SA) and jasmonic acid (JA). PLoS ONE 2015, 10, e0140609. [Google Scholar] [CrossRef] [PubMed]
  82. Gao, Z.; Meng, C.; Zhang, X.; Xu, D.; Miao, X.; Wang, Y.; Yang, L.; Lv, H.; Chen, L.; Ye, N. Induction of salicylic acid (SA) on transcriptional expression of eight carotenoid genes and astaxanthin accumulation in Haematococcus pluvialis. Enzym. Microb. Technol. 2012, 51, 225–230. [Google Scholar] [CrossRef] [PubMed]
  83. Du, F.; Hu, C.; Sun, X.; Xu, N. Transcriptome analysis reveals pathways responsible for the promoting effect of sucrose on astaxanthin accumulation in Haematococcus pluvialis under high light condition. Aquaculture 2021, 530, 735757. [Google Scholar] [CrossRef]
  84. Cheng, J.; Li, K.; Zhu, Y.; Yang, W.; Zhou, J.; Cen, K. Transcriptome sequencing and metabolic pathways of astaxanthin accumulated in Haematococcus pluvialis mutant under 15% CO2. Bioresour. Technol. 2017, 228, 99–105. [Google Scholar] [CrossRef]
  85. Hu, C.; Cui, D.; Sun, X.; Shi, J.; Song, L.; Li, Y.; Xu, N. Transcriptomic analysis unveils survival strategies of autotrophic Haematococcus pluvialis against high light stress. Aquaculture 2019, 513, 734430. [Google Scholar] [CrossRef]
  86. Che, H.; Li, Q.; Zhang, T.; Wang, D.; Yang, L.; Xu, J.; Yanagita, T.; Xue, C.; Chang, Y.; Wang, Y. Effects of astaxanthin and docosahexaenoic-acid-acylated astaxanthin on Alzheimer’s disease in APP/PS1 double-transgenic mice. J. Agric. Food. Chem. 2018, 66, 4948–4957. [Google Scholar] [CrossRef]
  87. Ma, R.; Thomas-Hall, S.R.; Chua, E.T.; Alsenani, F.; Eltanahy, E.; Netzel, M.E.; Netzel, G.; Lu, Y.; Schenk, P.M. Gene expression profiling of astaxanthin and fatty acid pathways in Haematococcus pluvialis in response to different LED lighting conditions. Bioresour. Technol. 2018, 250, 591–602. [Google Scholar] [CrossRef] [Green Version]
  88. Gao, Z.; Meng, C.; Zhang, X.; Xu, D.; Zhao, Y.; Wang, Y.; Lv, H.; Yang, L.; Chen, L.; Ye, N. Differential expression of carotenogenic genes, associated changes on astaxanthin production and photosynthesis features induced by JA in H. pluvialis. PLoS ONE 2012, 7, e42243. [Google Scholar] [CrossRef]
  89. Zhao, Y.; Wang, H.-P.; Yu, C.; Ding, W.; Han, B.; Geng, S.; Ning, D.; Ma, T.; Yu, X. Integration of physiological and metabolomic profiles to elucidate the regulatory mechanisms underlying the stimulatory effect of melatonin on astaxanthin and lipids coproduction in Haematococcus pluvialis under inductive stress conditions. Bioresour. Technol. 2021, 319, 124150. [Google Scholar] [CrossRef] [PubMed]
  90. Zhao, Y.; Cui, J.; Li, Q.; Qiao, T.; Zhong, D.-b.; Zhao, P.; Yu, X. A joint strategy comprising melatonin and 3-methyladenine to concurrently stimulate biomass and astaxanthin hyperaccumulation by Haematococcus pluvialis. Bioresour. Technol. 2021, 341, 125784. [Google Scholar] [CrossRef] [PubMed]
  91. He, B.; Hou, L.; Dong, M.; Shi, J.; Huang, X.; Ding, Y.; Cong, X.; Zhang, F.; Zhang, X.; Zang, X. Transcriptome analysis in Haematococcus pluvialis: Astaxanthin induction by high light with acetate and Fe2+. Int. J. Mol. Sci. 2018, 19, 175. [Google Scholar] [CrossRef] [Green Version]
  92. Lee, C.; Ahn, J.-W.; Kim, J.-B.; Kim, J.Y.; Choi, Y.-E. Comparative transcriptome analysis of Haematococcus pluvialis on astaxanthin biosynthesis in response to irradiation with red or blue LED wavelength. World J. Microbiol. Biotechnol. 2018, 34, 1–14. [Google Scholar] [CrossRef]
  93. Lu, Z.; Dai, J.; Zheng, L.; Teng, Z.; Zhang, Q.; Qiu, D.; Song, L. Disodium 2-oxoglutarate promotes carbon flux into astaxanthin and fatty acid biosynthesis pathways in Haematococcus. Bioresour. Technol. 2020, 299, 122612. [Google Scholar] [CrossRef]
  94. Wang, X.; Meng, C.; Zhang, H.; Xing, W.; Cao, K.; Zhu, B.; Zhang, C.; Sun, F.; Gao, Z. Transcriptomic and proteomic characterizations of the molecular response to blue light and salicylic acid in Haematococcus pluvialis. Mar. Drugs 2021, 20, 1. [Google Scholar] [CrossRef] [PubMed]
  95. Li, Y.; Cui, D.; Zhuo, P.; Zhang, L.; Sun, X.; Xu, N. A new approach to promote astaxanthin accumulation via Na2WO4 in Haematococcus pluvialis. J. Oceanol. Limnol. 2019, 37, 176–185. [Google Scholar] [CrossRef]
  96. Xu, R.; Zhang, L.; Yu, W.; Liu, J. A strategy for interfering with the formation of thick cell walls in Haematococcus pluvialis by down-regulating the mannan synthesis pathway. Bioresour. Technol. 2022, 362, 127783. [Google Scholar] [CrossRef]
  97. Wang, Q.; Oshita, K.; Takaoka, M.; Shiota, K. Influence of water content and cell disruption on lipid extraction using subcritical dimethyl ether in wet microalgae. Bioresour. Technol. 2021, 329, 124892. [Google Scholar] [CrossRef]
  98. Talebi, S.; Edalatpour, A.; Tavakoli, O. Algal biorefinery: A potential solution to the food–energy–water–environment nexus. Sustain. Energy Fuels 2022, 6, 2623–2664. [Google Scholar] [CrossRef]
  99. Nitsos, C.; Filali, R.; Taidi, B.; Lemaire, J. Current and novel approaches to downstream processing of microalgae: A review. Biotechnol. Adv. 2020, 45, 107650. [Google Scholar] [CrossRef]
  100. Ahirwar, A.; Meignen, G.; Khan, M.J.; Khan, N.; Rai, A.; Schoefs, B.; Marchand, J.; Varjani, S.; Vinayak, V. Nanotechnological approaches to disrupt the rigid cell walled microalgae grown in wastewater for value-added biocompounds: Commercial applications, challenges, and breakthrough. Biomass Convers. Biorefinery 2021, 1–26. [Google Scholar] [CrossRef]
  101. Ahirwar, A.; Khan, M.J.; Sirotiya, V.; Mourya, M.; Rai, A.; Schoefs, B.; Marchand, J.; Varjani, S.; Vinayak, V. Pulsed Electric Field–Assisted Cell Permeabilization of Microalgae (Haematococcus pluvialis) for Milking of Value-Added Compounds. BioEnergy Res. 2022, 16, 1–14. [Google Scholar] [CrossRef]
  102. Yao, J.; Kim, H.S.; Kim, J.Y.; Choi, Y.-E.; Park, J. Mechanical stress induced astaxanthin accumulation of H. pluvialis on a chip. Lab A Chip 2020, 20, 647–654. [Google Scholar] [CrossRef] [PubMed]
  103. Khan, M.J.; Bawra, N.; Verma, A.; Kumar, V.; Pugazhendhi, A.; Joshi, K.B.; Vinayak, V. Cultivation of diatom Pinnularia saprophila for lipid production: A comparison of methods for harvesting the lipid from the cells. Bioresour. Technol. 2021, 319, 124129. [Google Scholar] [CrossRef] [PubMed]
  104. Lee, N.; Narasimhan, A.L.; Moon, G.; Kim, Y.-E.; Park, M.; Kim, B.; Mahadi, R.; Chung, S.; Oh, Y.-K. Room-Temperature Cell Disruption and Astaxanthin Recovery from Haematococcus lacustris Cysts Using Ultrathin α-Quartz Nanoplates and Ionic Liquids. Appl. Sci. 2022, 12, 2210. [Google Scholar] [CrossRef]
  105. Praveenkumar, R.; Gwak, R.; Kang, M.; Shim, T.S.; Cho, S.; Lee, J.; Oh, Y.-K.; Lee, K.; Kim, B. Regenerative astaxanthin extraction from a single microalgal (Haematococcus pluvialis) cell using a gold nano-scalpel. ACS Appl. Mater. Interfaces 2015, 7, 22702–22708. [Google Scholar] [CrossRef] [PubMed]
  106. Song, K.; Li, Y.; Zhang, H.; An, N.; Wei, Y.; Wang, L.; Tian, C.; Yuan, M.; Sun, Y.; Xing, Y. Oxidative stress-mediated blood-brain barrier (BBB) disruption in neurological diseases. Oxidative Med. Cell. Longev. 2020, 2020, 1–27. [Google Scholar] [CrossRef]
  107. Reboul, E. Mechanisms of carotenoid intestinal absorption: Where do we stand? Nutrients 2019, 11, 838. [Google Scholar] [CrossRef] [Green Version]
  108. Kidd, P. Astaxanthin, cell membrane nutrient with diverse clinical benefits and anti-aging potential. Altern Med Rev 2011, 16, 355–364. [Google Scholar]
  109. Okada, Y.; Ishikura, M.; Maoka, T. Bioavailability of astaxanthin in Haematococcus algal extract: The effects of timing of diet and smoking habits. Biosci. Biotechnol. Biochem. 2009, 73, 1928–1932. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Chen, X.; Chen, R.; Guo, Z.; Li, C.; Li, P. The preparation and stability of the inclusion complex of astaxanthin with β-cyclodextrin. Food Chem. 2007, 101, 1580–1584. [Google Scholar] [CrossRef]
  111. Martínez-Álvarez, Ó.; Calvo, M.M.; Gómez-Estaca, J. Recent advances in astaxanthin micro/nanoencapsulation to improve its stability and functionality as a food ingredient. Mar. Drugs 2020, 18, 406. [Google Scholar] [CrossRef]
  112. Edelman, R.; Engelberg, S.; Fahoum, L.; Meyron-Holtz, E.G.; Livney, Y.D. Potato protein-based carriers for enhancing bioavailability of astaxanthin. Food Hydrocoll. 2019, 96, 72–80. [Google Scholar] [CrossRef]
  113. Li, C.; Song, Q.; Yin, X.; Song, R.; Chen, G. Preparation, Characterization, and In Vitro Anticancer Activity Evaluation of Broccoli-Derived Extracellular Vesicle-Coated Astaxanthin Nanoparticles. Molecules 2022, 27, 3955. [Google Scholar] [CrossRef]
  114. Jiang, G.-L.; Zhu, M.-J. Preparation of astaxanthin-encapsulated complex with zein and oligochitosan and its application in food processing. Lwt 2019, 106, 179–185. [Google Scholar] [CrossRef]
  115. Pan, L.; Wang, H.; Gu, K. Nanoliposomes as vehicles for astaxanthin: Characterization, in vitro release evaluation and structure. Molecules 2018, 23, 2822. [Google Scholar] [CrossRef] [Green Version]
  116. Liu, Y.; Huang, L.; Li, D.; Wang, Y.; Chen, Z.; Zou, C.; Liu, W.; Ma, Y.; Cao, M.-J.; Liu, G.-M. Re-assembled oleic acid-protein complexes as nano-vehicles for astaxanthin: Multispectral analysis and molecular docking. Food Hydrocoll. 2020, 103, 105689. [Google Scholar] [CrossRef]
  117. Hu, Q.; Hu, S.; Fleming, E.; Lee, J.-Y.; Luo, Y. Chitosan-caseinate-dextran ternary complex nanoparticles for potential oral delivery of astaxanthin with significantly improved bioactivity. Int. J. Biol. Macromol. 2020, 151, 747–756. [Google Scholar] [CrossRef]
  118. Zhang, X.; Yin, W.; Qi, Y.; Li, X.; Zhang, W.; He, G. Microencapsulation of astaxanthin in alginate using modified emulsion technology: Preparation, characterization, and cytostatic activity. Can. J. Chem. Eng. 2017, 95, 412–419. [Google Scholar] [CrossRef]
  119. Liu, X.; Xie, J.; Zhou, L.; Zhang, J.; Chen, Z.; Xiao, J.; Cao, Y.; Xiao, H. Recent advances in health benefits and bioavailability of dietary astaxanthin and its isomers. Food Chem. 2022, 404, 134605. [Google Scholar] [CrossRef] [PubMed]
  120. Liaqat, F.; Khazi, M.I.; Bahadar, A.; He, L.; Aslam, A.; Liaquat, R.; Agathos, S.N.; Li, J. Mixotrophic cultivation of microalgae for carotenoid production. Rev. Aquac. 2023, 15, 35–61. [Google Scholar] [CrossRef]
  121. Rao, A.R.; Sindhuja, H.; Dharmesh, S.M.; Sankar, K.U.; Sarada, R.; Ravishankar, G.A. Effective inhibition of skin cancer, tyrosinase, and antioxidative properties by astaxanthin and astaxanthin esters from the green alga Haematococcus pluvialis. J. Agric. Food Chem. 2013, 61, 3842–3851. [Google Scholar] [CrossRef] [PubMed]
  122. Visioli, F.; Artaria, C. Astaxanthin in cardiovascular health and disease: Mechanisms of action, therapeutic merits, and knowledge gaps. Food Funct. 2017, 8, 39–63. [Google Scholar] [CrossRef]
  123. Wu, D.; Xu, H.; Chen, J.; Zhang, L. Effects of astaxanthin supplementation on oxidative stress. Int. J. Vitam. Nutr. Res. 2019, 90, 179–194. [Google Scholar] [CrossRef] [PubMed]
  124. Conti, P.; Ronconi, G.; Caraffa, A.; Gallenga, C.; Ross, R.; Frydas, I.; Kritas, S. Induction of pro-inflammatory cytokines (IL-1 and IL-6) and lung inflammation by Coronavirus-19 (COVI-19 or SARS-CoV-2): Anti-inflammatory strategies. J. Biol. Regul Homeost Agents 2020, 34, 327–331. [Google Scholar]
  125. Zhang, C.; Wu, Z.; Li, J.-W.; Zhao, H.; Wang, G.-Q. Cytokine release syndrome in severe COVID-19: Interleukin-6 receptor antagonist tocilizumab may be the key to reduce mortality. Int. J. Antimicrob. Agents 2020, 55, 105954. [Google Scholar] [CrossRef]
  126. Shah, M.M.R.; Liang, Y.; Cheng, J.J.; Daroch, M. Astaxanthin-producing green microalga Haematococcus pluvialis: From single cell to high value commercial products. Front. Plant Sci. 2016, 7, 531. [Google Scholar] [CrossRef] [Green Version]
  127. Fassett, R.G.; Coombes, J.S. Astaxanthin: A potential therapeutic agent in cardiovascular disease. Mar. Drugs 2011, 9, 447–465. [Google Scholar] [CrossRef] [Green Version]
  128. Miyachi, M.; Matsuno, T.; Asano, K.; Mataga, I. Anti-inflammatory effects of astaxanthin in the human gingival keratinocyte line NDUSD-1. J. Clin. Biochem. Nutr. 2015, 56, 171–178. [Google Scholar] [CrossRef] [Green Version]
  129. Dhinaut, J.; Balourdet, A.; Teixeira, M.; Chogne, M.; Moret, Y. A dietary carotenoid reduces immunopathology and enhances longevity through an immune depressive effect in an insect model. Sci. Rep. 2017, 7, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Palozza, P.; Krinsky, N.I. Astaxanthin and canthaxanthin are potent antioxidants in a membrane model. Arch. Biochem. Biophys. 1992, 297, 291–295. [Google Scholar] [CrossRef]
  131. Ambati, R.R.; Siew Moi, P.; Ravi, S.; Aswathanarayana, R.G. Astaxanthin: Sources, extraction, stability, biological activities and its commercial applications—A review. Mar. Drugs 2014, 12, 128–152. [Google Scholar] [CrossRef] [PubMed]
  132. Liu, X.; Osawa, T. Cis astaxanthin and especially 9-cis astaxanthin exhibits a higher antioxidant activity in vitro compared to the all-trans isomer. Biochem. Biophys. Res. Commun. 2007, 357, 187–193. [Google Scholar] [CrossRef] [PubMed]
  133. Park, J.S.; Chyun, J.H.; Kim, Y.K.; Line, L.L.; Chew, B.P. Astaxanthin decreased oxidative stress and inflammation and enhanced immune response in humans. Nutr. Metab. 2010, 7, 1–10. [Google Scholar] [CrossRef] [Green Version]
  134. Fakhri, S.; Yosifova Aneva, I.; Farzaei, M.H.; Sobarzo-Sánchez, E. The Neuroprotective Effects of Astaxanthin: Therapeutic Targets and Clinical Perspective. Molecules 2019, 24, 2640. [Google Scholar] [CrossRef] [Green Version]
  135. Chang, M.X.; Xiong, F. Astaxanthin and its effects in inflammatory responses and inflammation-associated diseases: Recent advances and future directions. Molecules 2020, 25, 5342. [Google Scholar] [CrossRef]
  136. Fakhri, S.; Nouri, Z.; Moradi, S.Z.; Farzaei, M.H. Astaxanthin, COVID-19 and immune response: Focus on oxidative stress, apoptosis and autophagy. Phytother. Res. 2020, 34, 2790. [Google Scholar] [CrossRef]
  137. Kammeyer, A.; Luiten, R. Oxidation events and skin aging. Ageing Res. Rev. 2015, 21, 16–29. [Google Scholar] [CrossRef]
  138. Honda, M.; Murakami, K.; Osawa, Y.; Kawashima, Y.; Hirasawa, K.; Kuroda, I. Z-Isomers of astaxanthin exhibit greater bioavailability and tissue accumulation efficiency than the all-E-isomer. J. Agric. Food Chem. 2021, 69, 3489–3495. [Google Scholar] [CrossRef]
  139. Cao, Y.; Yang, L.; Qiao, X.; Xue, C.; Xu, J. Dietary astaxanthin: An excellent carotenoid with multiple health benefits. Crit. Rev. Food Sci. Nutr. 2021, 1–27. [Google Scholar] [CrossRef] [PubMed]
  140. Ozawa, Y. Oxidative stress in the light-exposed retina and its implication in age-related macular degeneration. Redox Biol. 2020, 37, 101779. [Google Scholar] [CrossRef] [PubMed]
  141. Wang, S.; Qi, X. The putative role of astaxanthin in neuroinflammation modulation: Mechanisms and therapeutic potential. Front. Pharmacol. 2022, 13, 916653. [Google Scholar] [CrossRef]
  142. Harada, F.; Morikawa, T.; Lennikov, A.; Mukwaya, A.; Schaupper, M.; Uehara, O.; Takai, R.; Yoshida, K.; Sato, J.; Horie, Y. Protective effects of oral astaxanthin nanopowder against ultraviolet-induced photokeratitis in mice. Oxidative Med. Cell. Longev. 2017, 2017, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Yasunori, N.; Miharu, M.; Jiro, T.; Akitoshi, K.; Yoshiharu, H.; Yuri, S.; Hiroki, T. The effect of astaxanthin on retinal capillary blood flow in normal volunteers. J. Clin. Ther. Med 2005, 21, 537–542. [Google Scholar]
  144. Karppi, J.; Rissanen, T.H.; Nyyssonen, K.; Kaikkonen, J.; Olsson, A.G.; Voutilainen, S.; Salonen, J.T. Effects of astaxanthin supplementation on lipid peroxidation. Int. J. Vitam. Nutr. Res. 2007, 77, 3–12. [Google Scholar] [CrossRef] [PubMed]
  145. Sorrenti, V.; Davinelli, S.; Scapagnini, G.; Willcox, B.J.; Allsopp, R.C.; Willcox, D.C. Astaxanthin as a putative geroprotector: Molecular basis and focus on brain aging. Mar. Drugs 2020, 18, 351. [Google Scholar] [CrossRef]
  146. Liu, N.; Chen, J.; Gao, D.; Li, W.; Zheng, D. Astaxanthin attenuates contrast agent-induced acute kidney injury in vitro and in vivo via the regulation of SIRT1/FOXO3a expression. Int. Urol. Nephrol. 2018, 50, 1171–1180. [Google Scholar] [CrossRef]
  147. Zhou, X.; Zhang, F.; Hu, X.; Chen, J.; Wen, X.; Sun, Y.; Liu, Y.; Tang, R.; Zheng, K.; Song, Y. Inhibition of inflammation by astaxanthin alleviates cognition deficits in diabetic mice. Physiol. Behav. 2015, 151, 412–420. [Google Scholar] [CrossRef]
  148. Ohgami, K.; Shiratori, K.; Kotake, S.; Nishida, T.; Mizuki, N.; Yazawa, K.; Ohno, S. Effects of astaxanthin on lipopolysaccharide-induced inflammation in vitro and in vivo. Investig. Ophthalmol. Vis. Sci. 2003, 44, 2694–2701. [Google Scholar] [CrossRef] [Green Version]
  149. Solomonov, Y.; Hadad, N.; Levy, R. The combined anti-inflammatory effect of astaxanthin, lyc-O-mato and carnosic acid in vitro and in vivo in a mouse model of peritonitis. J. Nutr. Food Sci. 2018, 8, 1000653. [Google Scholar] [CrossRef]
  150. Liu, G.; Shi, Y.; Peng, X.; Liu, H.; Peng, Y.; He, L. Astaxanthin attenuates adriamycin-induced focal segmental glomerulosclerosis. Pharmacology 2015, 95, 193–200. [Google Scholar] [CrossRef]
  151. Qiu, X.; Fu, K.; Zhao, X.; Zhang, Y.; Yuan, Y.; Zhang, S.; Gu, X.; Guo, H. Protective effects of astaxanthin against ischemia/reperfusion induced renal injury in mice. J. Transl. Med. 2015, 13, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Davinelli, S.; Melvang, H.M.; Andersen, L.P.; Scapagnini, G.; Nielsen, M.E. Astaxanthin from shrimp cephalothorax stimulates the immune response by enhancing IFN-γ, IL-10, and IL-2 secretion in splenocytes of Helicobacter pylori-infected mice. Mar. Drugs 2019, 17, 382. [Google Scholar] [CrossRef] [Green Version]
  153. Park, J.H.; Yeo, I.J.; Han, J.H.; Suh, J.W.; Lee, H.P.; Hong, J.T. Anti-inflammatory effect of astaxanthin in phthalic anhydride-induced atopic dermatitis animal model. Exp. Dermatol. 2018, 27, 378–385. [Google Scholar] [CrossRef] [PubMed]
  154. Bennedsen, M.; Wang, X.; Willén, R.; Wadström, T.; Andersen, L.P. Treatment of H. pylori infected mice with antioxidant astaxanthin reduces gastric inflammation, bacterial load and modulates cytokine release by splenocytes. Immunol. Lett. 2000, 70, 185–189. [Google Scholar] [CrossRef]
  155. Li, H.; Li, J.; Hou, C.; Li, J.; Peng, H.; Wang, Q. The effect of astaxanthin on inflammation in hyperosmolarity of experimental dry eye model in vitro and in vivo. Exp. Eye Res. 2020, 197, 108113. [Google Scholar] [CrossRef] [PubMed]
  156. Nakagawa, K.; Kiko, T.; Miyazawa, T.; Burdeos, G.C.; Kimura, F.; Satoh, A.; Miyazawa, T. Antioxidant effect of astaxanthin on phospholipid peroxidation in human erythrocytes. Br. J. Nutr. 2011, 105, 1563–1571. [Google Scholar] [CrossRef] [Green Version]
  157. Ye, Q.; Huang, B.; Zhang, X.; Zhu, Y.; Chen, X. Astaxanthin protects against MPP+-induced oxidative stress in PC12 cells via the HO-1/NOX2 axis. BMC Neurosci. 2012, 13, 1–13. [Google Scholar] [CrossRef] [Green Version]
  158. Haider, S.; Saleem, S.; Perveen, T.; Tabassum, S.; Batool, Z.; Sadir, S.; Liaquat, L.; Madiha, S. Age-related learning and memory deficits in rats: Role of altered brain neurotransmitters, acetylcholinesterase activity and changes in antioxidant defense system. Age 2014, 36, 1291–1302. [Google Scholar] [CrossRef]
  159. Al-Amin, M.M.; Akhter, S.; Hasan, A.T.; Alam, T.; Nageeb Hasan, S.; Saifullah, A.; Shohel, M. The antioxidant effect of astaxanthin is higher in young mice than aged: A region specific study on brain. Metab. Brain Dis. 2015, 30, 1237–1246. [Google Scholar] [CrossRef] [PubMed]
  160. Chang, C.-H.; Chen, C.-Y.; Chiou, J.-Y.; Peng, R.Y.; Peng, C.-H. Astaxanthine secured apoptotic death of PC12 cells induced by β-amyloid peptide 25–35: Its molecular action targets. J. Med. Food 2010, 13, 548–556. [Google Scholar] [CrossRef] [PubMed]
  161. Grimmig, B.; Daly, L.; Hudson, C.; Nash, K.; Bickford, P. Astaxanthin attenuates neurotoxicity in a mouse model of Parkinson’s disease. Funct. Foods Health Dis. 2017, 7, 562–576. [Google Scholar] [CrossRef]
  162. Kim, H.A.; Miller, A.A.; Drummond, G.R.; Thrift, A.G.; Arumugam, T.V.; Phan, T.G.; Srikanth, V.K.; Sobey, C.G. Vascular cognitive impairment and Alzheimer’s disease: Role of cerebral hypoperfusion and oxidative stress. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2012, 385, 953–959. [Google Scholar] [CrossRef]
  163. Lennikov, A.; Kitaichi, N.; Fukase, R.; Murata, M.; Noda, K.; Ando, R.; Ohguchi, T.; Kawakita, T.; Ohno, S.; Ishida, S. Amelioration of ultraviolet-induced photokeratitis in mice treated with astaxanthin eye drops. Mol. Vis. 2012, 18, 455. [Google Scholar] [PubMed]
  164. Dong, L.-Y.; Jin, J.; Lu, G.; Kang, X.-L. Astaxanthin attenuates the apoptosis of retinal ganglion cells in db/db mice by inhibition of oxidative stress. Mar. Drugs 2013, 11, 960–974. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Janani, R.; Anitha, R.E.; Perumal, M.K.; Divya, P.; Baskaran, V. Astaxanthin mediated regulation of VEGF through HIF1α and XBP1 signaling pathway: An insight from ARPE-19 cell and streptozotocin mediated diabetic rat model. Exp. Eye Res. 2021, 206, 108555. [Google Scholar] [CrossRef]
  166. Yeh, P.-T.; Huang, H.-W.; Yang, C.-M.; Yang, W.-S.; Yang, C.-H. Astaxanthin inhibits expression of retinal oxidative stress and inflammatory mediators in streptozotocin-induced diabetic rats. PLoS ONE 2016, 11, e0146438. [Google Scholar] [CrossRef] [Green Version]
  167. Kuraji, M.; Matsuno, T.; Satoh, T. Astaxanthin affects oxidative stress and hyposalivation in aging mice. J. Clin. Biochem. Nutr. 2016, 59, 79–85. [Google Scholar] [CrossRef] [Green Version]
  168. Xue, Y.; Qu, Z.; Fu, J.; Zhen, J.; Wang, W.; Cai, Y.; Wang, W. The protective effect of astaxanthin on learning and memory deficits and oxidative stress in a mouse model of repeated cerebral ischemia/reperfusion. Brain Res. Bull. 2017, 131, 221–228. [Google Scholar] [CrossRef]
  169. Lee, D.-H.; Lee, Y.J.; Kwon, K.H. Neuroprotective effects of astaxanthin in oxygen-glucose deprivation in SH-SY5Y cells and global cerebral ischemia in rat. J. Clin. Biochem. Nutr. 2010, 47, 121–129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Taksima, T.; Chonpathompikunlert, P.; Sroyraya, M.; Hutamekalin, P.; Limpawattana, M.; Klaypradit, W. Effects of astaxanthin from shrimp shell on oxidative stress and behavior in animal model of Alzheimer’s disease. Mar. Drugs 2019, 17, 628. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Waldvogel, F.A. Infectious diseases in the 21st century: Old challenges and new opportunities. Int. J. Infect. Dis. 2004, 8, 5–12. [Google Scholar] [CrossRef] [Green Version]
  172. Mohammed Ali, H.S.; Altayb, H.N.; Bayoumi, A.A.M.; El Omri, A.; Firoz, A.; Chaieb, K. In silico screening of the effectiveness of natural compounds from algae as SARS-CoV-2 inhibitors: Molecular docking, ADMT profile and molecular dynamic studies. J. Biomol. Struct. Dyn. 2022, 1–16. [Google Scholar] [CrossRef] [PubMed]
  173. Ahmadi, A.-R.; Ayazi-Nasrabadi, R. Astaxanthin protective barrier and its ability to improve the health in patients with COVID-19. Iran. J. Microbiol. 2021, 13, 434. [Google Scholar] [CrossRef] [PubMed]
  174. Honda, M.; Kageyama, H.; Hibino, T.; Osawa, Y.; Kawashima, Y.; Hirasawa, K.; Kuroda, I. Evaluation and improvement of storage stability of astaxanthin isomers in oils and fats. Food Chem. 2021, 352, 129371. [Google Scholar] [CrossRef]
Figure 1. Structure of different carotenoids.
Figure 1. Structure of different carotenoids.
Marinedrugs 21 00176 g001
Figure 2. Different isomers of astaxanthin.
Figure 2. Different isomers of astaxanthin.
Marinedrugs 21 00176 g002
Figure 3. Possibilities of chemical synthesis of astaxanthin by (I) Hydroxylation; (II) Oxidation; (III) Wittig reaction.
Figure 3. Possibilities of chemical synthesis of astaxanthin by (I) Hydroxylation; (II) Oxidation; (III) Wittig reaction.
Marinedrugs 21 00176 g003
Figure 4. Biosynthetic pathway of astaxanthin synthesis from Haematococcus. Reproduced with permissions from [17].
Figure 4. Biosynthetic pathway of astaxanthin synthesis from Haematococcus. Reproduced with permissions from [17].
Marinedrugs 21 00176 g004
Figure 5. Economical cultivation of Haematococcus lacustris for astaxanthin: (I) Photosynthetic microalgal microbial fuel cell with Haematococcus lacustris and cells showing their harvest to astaxanthin on being applied optimized high pressure/pulse electric current. (II) Bubble farming in a plastic bubble wrap showing Haematococcus lacustris cells being grown in it with no water loss and sufficient exchange of gases.
Figure 5. Economical cultivation of Haematococcus lacustris for astaxanthin: (I) Photosynthetic microalgal microbial fuel cell with Haematococcus lacustris and cells showing their harvest to astaxanthin on being applied optimized high pressure/pulse electric current. (II) Bubble farming in a plastic bubble wrap showing Haematococcus lacustris cells being grown in it with no water loss and sufficient exchange of gases.
Marinedrugs 21 00176 g005
Figure 6. Harvesting techniques showing better astaxanthin synthesis, accumulation without cell lysis, and its bioavailability, where I: Intact Haematococcus rich in astaxanthin.; Cell wall fracturing by IIA: Electrolysis; IIB: Nanoparticles; IIC: Mechanical shear strain; IID: Gold nano scalpels; III: Haematocccus cell oozing astaxanthin without cell lysis; IV: Rejuvenation and healing of Haematococcus cells for more asataxanthin and V: Collection of soluble astaxanthin by encapsulation.
Figure 6. Harvesting techniques showing better astaxanthin synthesis, accumulation without cell lysis, and its bioavailability, where I: Intact Haematococcus rich in astaxanthin.; Cell wall fracturing by IIA: Electrolysis; IIB: Nanoparticles; IIC: Mechanical shear strain; IID: Gold nano scalpels; III: Haematocccus cell oozing astaxanthin without cell lysis; IV: Rejuvenation and healing of Haematococcus cells for more asataxanthin and V: Collection of soluble astaxanthin by encapsulation.
Marinedrugs 21 00176 g006
Figure 7. Different mechanism of astaxanthin action inside human body. Note: Nrf2: Nuclear factor erythroid 2-related factor 2; Keap: Kelch-like ECH-associated protein 1; HO-1: Heme oxygenase-1 (HO-1); BAX: Bcl-2 Associated X-protein; IL6: Interleukin 6; TNF-ALPHA: Tumour Necrosis Factor alpha; ICA: intercellular adhesion molecule 1; IKB: inhibitor of nuclear factor kappa B; ARE: Antioxidant response element; NQ01: NAD(P)H quinone dehydrogenase 1.
Figure 7. Different mechanism of astaxanthin action inside human body. Note: Nrf2: Nuclear factor erythroid 2-related factor 2; Keap: Kelch-like ECH-associated protein 1; HO-1: Heme oxygenase-1 (HO-1); BAX: Bcl-2 Associated X-protein; IL6: Interleukin 6; TNF-ALPHA: Tumour Necrosis Factor alpha; ICA: intercellular adhesion molecule 1; IKB: inhibitor of nuclear factor kappa B; ARE: Antioxidant response element; NQ01: NAD(P)H quinone dehydrogenase 1.
Marinedrugs 21 00176 g007
Table 1. Chemicals/factors induced astaxanthin biosynthesis and accumulation by the upregulation of the carotenogenic genes in Haematococcus lacustris.
Table 1. Chemicals/factors induced astaxanthin biosynthesis and accumulation by the upregulation of the carotenogenic genes in Haematococcus lacustris.
Factors/ChemicalsAstaxanthin Biosynthesis and Gene ExpressionReference
Salicylic acid and jasmonic acidUp-regulation of carotenogenic genes of psy, pds, zds and crtR-B[81]
Salicylic acidUp-regulation by ipi-1, ipi-2, psy, crtR-B, bkt and crtO genes at transcriptional level.
lyc gene at post-transcriptional level and pds gene at both.
[82]
Potassium iodide and high lightUpregulation of transcription levels of 15-cis-phytoene/all-transphytoene synthase gene (CrtB) and 15-cis-phytoene desaturase gene (PDS).[78]
Trisodium citrateUp-regulation and the expression of genes encoding beta-ring hydroxylase (LUT5), beta-carotene/zeaxanthin 4-ketolase (CrtW) and beta-carotene 3-hydroxylase (CrtZ).[83]
CO2Upregulation of genes involved in β-carotene biosynthesis such as PSY, ZDS and lcyB and β-carotene conversion like crtZ.[84]
High light stressUp-regulation and the expression of genes in MEP and astaxanthin biosynthesis pathway genes (ISPF, GGPS, PDS, CrtW and CrtZ)
and concurrent down-regulation the expression of SPS, CHLP and CrtL-e.
[85]
Low temperature plasmaExpression of IPT9 and CYP735A1 involved in zeatin synthesis, amidase gene (AMI1), aldehyde dehydrogenase gene (ALDH7A1) in the indole-3-acetic acid (IAA) synthesis pathway, acyl-CoA oxidase gene (ACX1) in methyl Jasmonates synthesis pathway[86]
High intensity blue and white LED lightUnder white light upregulation of transcripts of astaxanthin biosynthesis genes psy, crtO, and bkt2.
and under blue light upregulation of genes psy, lcy, crtO, and crtR-B.
[87]
Jasmonic acidAstaxanthin biosynthesis up-regulation by psy, pds, crtR-B, lyc, bkt2 and crtO at the transcriptional level and ipi-1, ipi-2 at both transcriptional and post-transcriptional levels.[88]
Blue, white, and red lightBlue light receptor gene upregulation of the biosynthesis pathway genes psy and pds, as well as dgat1 and dgat2d.[87]
MelatoninUpregulation of the major metabolites of the TCA cycle and the GABA shunt.[89]
Melatonin + 3-methyladenineThe gene encoding lycopene β-cyclase (LCY), which catalyses lycopene to β-carotene which acts as the direct precursor for the accumulation of astaxanthin, which is further catalysed by β-carotene hydroxylase (CHY).[90]
Melatonin + CalciumUp-regulation and the expression levels of astaxanthin biosynthetic genes (lcy, lycopene β-cyclase; bkt, β-carotene ketolase; chy, β-carotene hydroxylase)[54]
Acetate and Fe2+PDS, crtISO, LUT1, LUT5, lcyB, lcyE, crtZ, CCD8, ZEP[91]
Under the monochromatic red (660 nm) or blue (450 nm) light-emitting diode (LED) irradiationUpregulation and expression of gene IPP (encoding isopentenyl pyrophosphate), PSY (encoding phytoene synthase), PDS (encoding phytoene desaturase), LYC (encoding lycopene beta cyclase), BKT (encoding beta-carotene ketolase), CHY (encoding carotenoid hydroxylase), and CBR (encoding carotene biosynthesis-related protein (chlorophyll a-b binding protein)[92]
Disodium 2-oxoglutarate (2-OG-2Na)Up-regulation of genes ipi, bkt and crtR-b.[93]
SucroseThe expression of genes encoding phytoene synthase (CrtB), beta-carotene/zeaxanthin 4-ketolase (CrtW, BKT) and beta-carotene 3- hydroxylase (CrtZ).[83]
Blue light and salicylic acidUnder blue light the expression of fpps was significantly regulated, and blue light reduced the expression of genes involved in astaxanthin synthesis.[94]
Na2 WO4Expression of ipi, psy and bkt for astaxanthin biosynthesis[95]
Note: PSY: Phytoene synthase; PDS: phytoene desaturase; ZDS: zeta-carotene desaturase; crtR-B: β-carotene hydroxylase; BKT: β-carotene ketolase; beta-ring hydroxylase (LUT5); ZDS: zeta-carotene desaturase; LcyB: lycopene beta-cyclase; MEP: methylerythritol 4-phosphate; PDS Phytoene desaturase; LYC: Lycopene beta cyclase; CHY: carotenoid hydroxylase; CBR: Carotene biosynthesis-related protein SPS: Sucrose phosphate synthase; CHLP: Chloroplastic; Lcy: Lycopene; TCA cycle: Tricarboxylic acid cycle; EMP pathway: Embden-Meyerhof-Parnas; PPP: Pentose phosphate pathway; IPP: Isopentenyl pyrophosphate; beta-carotene 3- hydroxylase (CrtZ); crtISO: Carotenoid isomerase; LUT1: Long undecoded transcript isoform-1; LUT5: Long undecoded transcript isoform-5; lcyB: Lycopene β cyclase; lcyE: Lycopene epsilon cyclase; crtZ: Beta-carotene hydroxylase; CCD8: Carotenoid cleavage dioxygenase 8; ZEP: Zeaxanthin epoxidase; ISPF: Isoprenoid F; GGPS: Geranylgeranyl pyrophosphate synthase.
Table 2. Therapeutic actions of astaxanthin at different concentration on different gene markers in different test animals.
Table 2. Therapeutic actions of astaxanthin at different concentration on different gene markers in different test animals.
Test
Parameter
ModelAstaxanthin ConcentrationTarget Gene/BiomarkerReference
Anti-inflammationMice25 mgkg−1 day−1NF-κB, TNF-α[147]
Rats1, 10 or 100 mgkg−1TNF-α, PGE2, IL-1βp-IKKα, p-IκBα[148]
BV-2 cells50 µMNF-κB p65, IL-6, MAPK[149]
Male Balb/c mice50 mgkg−1Nrf2, NLRP3, IL-1β, IL-18[150]
Male ICR mice5 mgkg−1 day−1IL-1β, IL-6, TNF-α[151]
BALB/c female mice10 or 40 mg d−1IL-2 and IL-10, IFN-γ[152]
HR-1 mice10 µg or 20 µgcm−2NF-κB, IL-1β, IL-6, TNF-α, IgE, COX-2, iNOS[153]
Balb/cA mice200 mg kg−1 body weight day−1IL-2, IFN-γ, IL-4[154]
BALB/c mice1 µL drop of 5 µMPI3K/Akt, HMGB1, TNF-α, IL-1β[155]
AntioxidantHuman6 or 12 mgd−1PLOOH[156]
PC12 cells5, 10, 20 µMNOX2, Sp1/NR, NFR2, HO-1[157]
Mice2 mg kg−1APOP, SOD, GSH, MDA[158]
PC12 cells0.1 µMNF-κB, Bax, IL-1β, TNFα[159]
Primary hippocampal neurons0.1 µMRyR2, NFATc4[160]
SH-SY5Y cells100 nMCYTc, PARP[161]
Motor neurons100 nMSOD1[162]
Ocular healthC57BL/6J mice1/0.1/ 0.01 ng ml−1 by eye dropNF-κB[163]
Db/db rats25/5 mg kg−1 (oral gavage)8-OHdG, SOD, MDA[164]
C57BL/6J mice50 mg.kg−1Bax, Bcl-2, Nrf2. HO1, ROS[165]
Wistar rats0.6/3 mg kg−1acrolein, 8-OHdG, NO, MCP-1, ICAM-1[166]
ddY mice100 mg kg−14-HNE, 8-OHdG[167]
NeuroprotectiveRats10 mg kg−1MVA, Nef2, SOD[168]
SH-SY5Y cells and Rats10 to 50 µM (cells) 30 mg kg−1 (rats)HSPs, iNOS[162]
Mice20 mg kg−1SOD, GHS, Casp3, Cyt C[169]
Alzheimer’s DiseaseWistar rats10 mg kg−1 body weightOxidative markers[170]
Note: NF-Κb: Nuclear factor kappa B subunit 1; TNF-α: Tumor necrosis factor alpha; PGE: Prostaglandin E synthase; IL-1βp-Interleukin (IL)-1β-; IKKα: inhibitor of nuclear factor kappa-B kinase subunit beta; p-IκBα: nuclear factor of kappa light polypeptide gene; IL-6: Interleukin-6; MAPK: Mitogen-activated protein kinase; NLRP3: Nucleotide binding domain, leucine rich containing family domain containing 3; IL-18: Interleukin-18; IL-2: Interleukin-2; IL-10: Interleukin-10; IFN-γ: Interferon gamma; IgE: Immunoglobulin E; COX-2: Cyclo-oxygenase 2; iNOS: Inducible nitric oxide synthase; PI3K/Akt: phosphoinositide-3-kinase–protein kinase B/Akt; HMGB1: High mobility group box protein 1; PLOOH: Phospholipid hydroperoxide; NOX2: phagocyte NADPH oxidase; Sp1/N: Transcription factor Sp1 N terminal; NFR2: Nuclear factor erythroid factor 2 related factor; HO-1: Heme oxygenase 1; APOP: Amyloid Beta Precursor Protein; SOD: Superoxide dismutase; GSH: Glutathion; MDA: Malondialdehyde; RyR2: Ryanodine receptor 2; NFATc4: Nuclear factor of activated T cells 4; CYTc: Cytochrome c; PARP: Poly(ADP- ribose) polymerase 1; SOD1: Superoxide dismutase 1; 8-OHdG: 8-hydroxydeoxyguanosine; NO: Nitric oxide; MCP-1: Monocyte chemoattractant protein-1; ICAM-1: Intercellular adhesion molecule 1; 4-HNE: 4-hydroxynonenal; MVA: Mevalonate pathway; Nef2: nuclear factor erythroid 2–related factor 2; Casp3: Caspase-3; Cyt C: Cytochrome C.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Nair, A.; Ahirwar, A.; Singh, S.; Lodhi, R.; Lodhi, A.; Rai, A.; Jadhav, D.A.; Harish; Varjani, S.; Singh, G.; et al. Astaxanthin as a King of Ketocarotenoids: Structure, Synthesis, Accumulation, Bioavailability and Antioxidant Properties. Mar. Drugs 2023, 21, 176. https://doi.org/10.3390/md21030176

AMA Style

Nair A, Ahirwar A, Singh S, Lodhi R, Lodhi A, Rai A, Jadhav DA, Harish, Varjani S, Singh G, et al. Astaxanthin as a King of Ketocarotenoids: Structure, Synthesis, Accumulation, Bioavailability and Antioxidant Properties. Marine Drugs. 2023; 21(3):176. https://doi.org/10.3390/md21030176

Chicago/Turabian Style

Nair, Anagha, Ankesh Ahirwar, Shashikala Singh, Reeta Lodhi, Aishwarya Lodhi, Anshuman Rai, Dipak A Jadhav, Harish, Sunita Varjani, Gurpreet Singh, and et al. 2023. "Astaxanthin as a King of Ketocarotenoids: Structure, Synthesis, Accumulation, Bioavailability and Antioxidant Properties" Marine Drugs 21, no. 3: 176. https://doi.org/10.3390/md21030176

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop